Skip to main content

Full text of "Organic and inorganic-organic thin film structures by molecular layer deposition: A review."

See other formats


BEILSTEIN JOURNAL OF NANOTECHNOLOGY 



Organic and inorganic-organic thin film structures by 
molecular layer deposition: A review 

Pia Sundberg and Maarit Karppinen*§ 



Review 



Address: 

Department of Chemistry, Aalto University, P.O. Box 16100 FI-00076 
Aalto, Finland 

Email: 

Maarit Karppinen* - maarit.karppinen@aalto.fi 

* Corresponding author 
§ FAX: +358 9 462 373 

Keywords: 

atomic layer deposition (ALD); hybrid inorganic-organic thin films; 
molecular layer deposition (MLD); nanolaminates; nanostructuring; 
organic thin films; superlattices; thin-film technology 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 
doi:10.3762/bjnano.5.123 

Received: 20 February 2014 
Accepted: 20 June 2014 
Published: 22 July 2014 

Associate Editor: A. Golzhauser 

© 2014 Sundberg and Karppinen; licensee Beilstein-lnstitut. 
License and terms: see end of document. 



Abstract 

The possibility to deposit purely organic and hybrid inorganic-organic materials in a way parallel to the state-of-the-art gas-phase 
deposition method of inorganic thin films, i.e., atomic layer deposition (ALD), is currently experiencing a strongly growing interest. 
Like ALD in case of the inorganics, the emerging molecular layer deposition (MLD) technique for organic constituents can be 
employed to fabricate high-quality thin films and coatings with thickness and composition control on the molecular scale, even on 
complex three-dimensional structures. Moreover, by combining the two techniques, ALD and MLD, fundamentally new types of 
inorganic-organic hybrid materials can be produced. In this review article, we first describe the basic concepts regarding the MLD 
and ALD/MLD processes, followed by a comprehensive review of the various precursors and precursor pairs so far employed in 
these processes. Finally, we discuss the first proof-of-concept experiments in which the newly developed MLD and ALD/MLD 
processes are exploited to fabricate novel multilayer and nanostructure architectures by combining different inorganic, organic and 
hybrid material layers into on-demand designed mixtures, superlattices and nanolaminates, and employing new innovative 
nanotemplates or post-deposition treatments to, e.g., selectively decompose parts of the structure. Such layer-engineered and/or 
nanostructured hybrid materials with exciting combinations of functional properties hold great promise for high-end technological 
applications. 



Introduction 

Many high-end technologies rely on our capability to fabricate 
thin films and coatings with on-demand tailored compositions 
and architectures in a highly controlled way. The atomic layer 
deposition (ALD) technique is capable of producing high- 



quality nanometer-scale thin films in an atomic layer-by-layer 
manner. Compared with other advanced gas-phase thin-film 
deposition techniques, ALD has several distinct advantages: 
The films can be deposited with a great control over the film 



1104 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 



thickness and they are not only pinhole free, dense and uniform, 
but also conformal even when deposited on complex three- 
dimensional (3D) structures. These features make ALD a 
method of choice for nanotechnology, for both material syn- 
thesis and device fabrication. The technology spectrum in 
which ALD can be utilized is extremely wide, including micro- 
electronics, catalysis, energy applications and protective/barrier 
coatings. 

The history of ALD goes back to the 1960s and 1970s [1-4]. 
Traditionally, ALD has been used to fabricate rather simple 
well-known inorganic materials, such as binary oxides and 
nitrides. The range of materials was fundamentally broadened 
by experiments producing organic polymers in the 1990s by a 
variant of ALD, now commonly known as molecular layer 
deposition (MLD), named after the molecular layer-by-layer 
fashion the film grows during the deposition [5-9], Then - most 
excitingly - in the late 2000s the two techniques, ALD and 
MLD, were combined to produce inorganic-organic hybrid ma- 
terials (Figure 1), making it possible to synthesize totally new 
material families with versatile characteristics, which are not 
accessible by any other existing technique [10-14]. 



■ organic 




■ hybrid 








II 1 1 1 1 1 1 1 li 


1 









1990 1995 2000 2005 2010 
Year 



Figure 1 : Number of articles annually published featuring organic and 
hybrid inorganic-organic thin films deposited by MLD and ALD/MLD. 

In the combined ALD/MLD process organic molecules are 
covalently bonded to the metal atoms and vice versa, forming 
periodic thin-film structures that can be imagined to consist of 
either interlinked hybrid inorganic-organic polymer chains of 
essentially identical lengths or alternating two-dimensional (2D) 
planes of inorganic and organic monolayers (Figure 2). The 
hybrid thin films may not only possess properties combined 
from those of the two parent materials, but may also have 
completely new material properties, making them excellent 
candidates for a wide range of applications. Possible uses for 



the hybrid ALD/MLD films include optoelectronic devices, 
sensors, flexible electronics, solar cell applications, and protec- 
tive coatings, to name only a few. It is also straightforward to 
make porous structures from the ALD/MLD grown hybrids by 
removing the organic part by simple annealing or wet-etching 
procedures [15,16]. Further tuning of material properties may 
be achieved by combining different inorganic, organic and 
hybrid layers into various thin-film mixtures, superstructures 
and nanolaminates. For example, precise control of the refrac- 
tive index is extremely important in optical applications [17], 
while control of the electrical properties is required for storage 
capacitors, non-volatile memories as well as for transparent 
thin- film transistors [18,19]. Moreover, the tunability of the 
surface roughness is advantageous when fabricating gas sensors 
[20]. 




Figure 2: Schematic illustration of purely organic thin films grown by 
MLD (left) and hybrid inorganic-organic thin films grown by ALD/MLD 
(right). 



Over the years a number of excellent reviews featuring various 
types of ALD processes have been published, most recently, 
e.g., by Puurunen [4], George [21] and Miikkulainen et al. [22]. 
A review by Knez et al. [23] focuses on nanostructure fabrica- 
tion by ALD. Although the introduction of the MLD method 
dates back two decades, the number of articles featuring purely 
organic thin films is still quite limited. Nevertheless some 
reviews concerning MLD-based thin films have been published 
in the past: George et al. [24] discuss the surface chemistry of 
MLD grown materials, addressing the problems which arise 
when using organic precursors in the growth process; Leskela et 
al. [25] shortly review the novel materials fabricated by ALD 
and MLD; George [26], George et al. [27] and Lee et al. [28] 



1105 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 



focus on metal alkoxide thin films; Yoshimura et al. [29] 
discuss a possibility to utilize MLD in cancer therapy applica- 
tions; King et al. [30] describe fine particle functionalization by 
ALD and MLD; and the review by Zhou et al. [31] covers all 
the organic interfaces fabricated by MLD. 

The aim of this review is to provide a thorough investigation of 
the various thin films deposited by taking advantage of the 
currently strongly emerging MLD technique, including pure 
organic thin films, hybrid inorganic-organic thin films and their 
mixtures and nanolaminate structures. First we will describe the 
sequential ALD/MLD process, followed by a few words about 
the organic precursors used in these processes. Then all the 
various materials fabricated utilizing MLD are reviewed: The 
purely organic materials are summarized first and the inor- 
ganic-organic materials are discussed in a separate chapter. 
Lastly, the variously mixed and nanostructured ALD/MLD and 
MLD materials are presented. 

Review 

Deposition cycle and ideal ALD/MLD growth 

In both ALD and MLD the gas-solid reactions occur in a self- 
limiting, surface-saturated manner. The characteristic ALD/ 
MLD growth can be described by a so-called ALD and/or MLD 
cycle. The number of precursors employed during an ALD or 
MLD process can be varied, but a prototype process is based on 
two. For example, in case of the hybrid inorganic-organic films 
one inorganic and one organic precursor are used, and the ALD/ 
MLD cycle can be separated into four steps consisting of 
precursor pulsing and intermediate purging steps as described in 
Figure 3. 

To ensure the self-limiting growth both the precursor pulsing 
and purging steps should be sufficiently long. In an ideal 
process the surface is fully covered with the precursor in each 
precursor pulsing step, but in practice only a partial coverage is 
typically achieved. The so-called growth-per-cycle (GPC) value 
is the average increase in film thickness during one ALD/MLD 
cycle. 

When the GPC remains constant with increasing number of 
deposition cycles, the growth is said to be linear. In some cases, 
however, the GPC is not constant from the beginning. The sub- 
strate may inhibit or enhance the film growth depending on the 
compatibility of chemistries of the substrate surface and the 
growing film, in which case the GPC is initially lower or higher 
before settling to a constant value [4]. It is the sequential self- 
limiting nature of ALD and MLD that enables the great thick- 
ness control and conformal growth of the films, which in turn 
makes the two techniques, ALD and MLD, and their combina- 
tions such a great asset for nanotechnology. 




Figure 3: An ALD/MLD cycle consisting of the following four steps: 
(1 ) the first (inorganic) precursor is pulsed to the reactor and it reacts 
with the surface species, (2) the excess precursor and possible 
byproducts are removed from the reactor, either by purging with inert 
gas such as nitrogen or argon, or by evacuation, (3) the second 
(organic) precursor is pulsed to the reactor and it reacts with the 
surface species, and finally (4) the excess precursor/possible byprod- 
ucts are removed from the reactor. In an ideal case a monolayer of a 
hybrid inorganic-organic material is formed. To deposit thicker films 
this basic ALD/MLD cycle is repeated as many times as needed to 
reach the targeted film thickness. 

The ALD or MLD growth typically depends on the deposition 
temperature at least in some temperature ranges. The effect of 
temperature on the GPC value is often described by a concept 
known as an ALD (or MLD) window (Figure 4a). The ALD 
window has been defined as a regime in which the GPC 
remains constant and does not depend on process parameters 
like temperature, gas pressure, precursor flows or purging 
times. Outside the ALD window the GPC value may be higher 
due to precursor condensation (at too low deposition tempera- 
tures) or decomposition (at too high deposition temperatures), 
whereas limited growth may result from insufficient reactivity 
(at too low deposition temperatures) or desorption (at too high 
deposition temperatures) of the precursor. However, the exis- 
tence of an ALD window is not a necessary prerequisite for an 
ALD-type growth, and such a window is not found for all well- 
behaving ALD (or MLD) processes. Examples of typical cases 
in which no temperature range of constant growth is seen, but 
the process may yet be highly reproducible are shown in 
Figure 4b-d. The growth may occur in the way shown in 
Figure 4b when the growth is not fully of ALD type, but one of 
the precursors diffuses into the film, improving the growth by 
providing more reactive sites: The diffusion out of the film is 
enhanced at higher temperatures, resulting in a lower growth 



1106 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 




Figure 4: Dependence of the film growth on the deposition temperature: (a) within the so-called ALD window the growth per cycle remains constant 
with increasing temperature, whereas (b)-(d) represent typical cases in which no temperature range of constant growth is seen, but the process may 
yet be highly reproducible. 



rate [12]. The decrease in growth at increasing deposition 
temperatures may be observed in general when the temperature 
affects the number of reactive sites or the reaction mechanism. 
The combined effect of reaction activation (increase in growth), 
followed by decrease on reactive sites may result in a growth 
such as shown in Figure 4d [4]. 

The MLD and ALD/MLD films often show lower than antici- 
pated GPC values. There are several possible causes for the 




M T I ? 

Figure 5: Ideally, the organic precursor molecule reacts with one 
surface site only and remains straight (left). It may also react twice with 
the surface (middle) or tilt (right). 



hindered growth (Figure 5). Organic precursor molecules with 
long chains are likely to tilt such that the growth is not perfectly 
perpendicular to the surface. Likewise, organic molecules may 
bend and react twice with the surface, reducing the number of 
reactive surface sites and lowering the growth rate. Organic 
precursors are also often bulky, causing steric hindrance. The 
various difficulties encountered when using organic precursors 
are discussed in detail in a review by George et al. [24]. Several 
strategies have been employed to improve the controllability of 
the growth process, such as using organic precursors with stiff 
backbones [13,14,32-37] or with two different functional 
groups [36-39], using reactions requiring surface activation 
[10,39-46], using precursors in which ring-opening reactions 
occur [47], or using three different precursors instead of two 
[48,49]. 

The ratio, r = GPCIML, where ML is the ideal length of the 
M-R monomer (without bending, see Figure 5), provides us 
with a measure of the perfectness of the growth. Thus, ideally 
r = 1. The r value achieved varies greatly with different 
precursor combinations. For purely organic MLD films r is 
typically lower than 0.5, exceptions being the hexanedioyl 
dichloride+hexane-l,6-diamine [8,50] and heptane- 1,7-di- 
amine+nonanedioyl dichloride [7] systems. For hybrid ALD/ 
MLD thin films there is a larger variation in the r values 



1107 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



depending on the organic precursor employed. The choice of 
the metal precursor seems to have a significant effect, too. For 
example, with the linear ethane- 1,2-diol (ethylene glycol, EG) 
molecule as the organic precursor and trimethylaluminum 
(A1(CH 3 ) 3 , TMA), diethylzinc (Zn(C 2 H 5 ) 2 , DEZ), titanium 
tetrachloride (TiC^) and zirconium ?er?-butoxide 
(Zr(CH4H90)4, ZTB) as the inorganic precursors the growth 
processes have yielded r values of 0.6 0.1, 0.6 and 0.2, respect- 
ively [12,51-54]. In case of the aromatic precursor benzene- 1,4- 
diol (hydroquinone, HQ), r values of 0.4 and 0.2 have been 
achieved for Al-based [13] and Zn-based [33] films, respective- 
ly. Systems exhibiting r values close to unity have been 
reported, such as TiCl4+4-aminophenol [37] and hexa-2,4- 
diyne-l,6-diol (HDD) [55,56] containing hybrid films. In case 
of the TiCl4+4-aminophenol process, the excellent growth 
could be attributed to the two different functional groups and 
the stiff aromatic backbone of 4-aminophenol as well as to the 
small -CI ligands in TiCl4 [37]. The HDD molecule with two 
triple bonds is also stiff and during the deposition process also 
the formation of bridging alkanes is induced by UV radiation 
after precursor pulsing steps [55,56]. However, as the hybrid 
systems seem to be sensitive regarding the process parameters, 
especially considering pulsing and purging times, there are 
systems with r values which are considerably higher than 1, 
e.g., TMA+heptanedioic acid [57] and TMA+oxiran-2- 
ylmethanol (glycidol, GLY) [38]. 

The characterization techniques used to investigate the thin 
films deposited by using MLD do not vary much from those 
techniques used for inorganic thin films grown by ALD. An in 
situ quartz crystal microbalance (QCM) is often used to give 



some insight on the growth dynamics of the deposition. Besides 
thickness measurements, X-ray reflectivity (XRR) can be used 
for evaluating densities and roughnesses of the thin films. The 
crystallinity of the films is examined by X-ray diffraction 
(XRD). The topography of the films can be investigated by 
using atomic force microscopy (AFM). Fourier transform 
infrared (FTIR) spectroscopy is useful for analyzing the chem- 
ical state of the films. The composition of the films can be 
studied by X-ray photoelectron spectroscopy (XPS), whereas 
the presence of a metal can be verified by X-ray fluorescence 
(XRF) measurements. Nanoindentation gives insight on the 
mechanical properties of the films. 

Organic precursors employed in MLD 

Both ALD and MLD set some requirements for the precursors 
employed, such as sufficient vapor pressure, reactivity and 
stability at the reaction temperature, to ensure feasible film 
growth. Finding organic compounds which would fulfill these 
requirements is not straightforward. Many of the organic 
precursors exhibit low vapor pressures at room temperature and 
it is thus mandatory to heat them to achieve a sufficient 
precursor supply. In Table 1, we list all the organic compounds 
employed/investigated as precursors for ALD/MLD. Here it 
should be noted that organic compounds typically have several 
different names; the nomenclature we use is based on the 
recommendations of the International Union of Pure and 
Applied Chemistry (IUPAC), but in Table 1 commonly used 
other names for the compounds are also given. It should also be 
emphasized that not all the processes based on the precursors 
listed in Table 1 exhibit the characteristic features of an ALD/ 
MLD process. In Table 1 we give - when accessible - the vapor 



Table 1 : Organic compounds (and their different names and abbreviations) employed in MLD and ALD/MLD processes together with vapor pres- 
sures, P, at 100 °C and temperatures, T, corresponding to a vapor pressure of 2 mbar for some of the organic precursors (the values were calculated 



IUPAC name 



abbreviation names used in references 



P(Pa) 7" CO 



(1£)-prop-1-ene-1,2,3-tricarboxylic acid 

(2£,4E)-hexa-2,4-dienedioic acid 

(2S)-2-aminopentanedioic acid 
(E)-butenedioic acid 
(Z)-butenedioic acid 

1,2-bis[(diamethylamino)dimethylsilyl]ethane 
1 ,4-diaminobenzene 

1 ,4-diisocyanatobenzene PDIC 

1 ,4-diisocyanatobutane 

1 ,4-diisothiocyanatobenzene 

2-aminoethanol 

2-oxepanone 



2,2'-(propane-2,2-diylbis(oxy))-diethanamine 

frans-aconitic acid 

(2E,4E)-hexa-2,4-dienedioic acid; 
frans.frans-muconic acid 

L-glutamic acid 

fumaric acid; (E)-butenedioic acid 

maleic acid; (Z)-butenedioic acid 

1,2-bis[(diamethylamino)dimethylsilyl]ethane 

p-phenylenediamine; 1 ,4-phenylenediamine 

1,4-phenylene diisocyanate 

1 ,4-diisocyanatobutane 

1,4-phenylene diisothiocyanate 

ethanolamine 

e-caprolactone 



6.37 



154 



6410 
1520 



182 
165 
142 

104 



43 
58 



1108 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



Table 1 : Organic compounds (and their different names and abbreviations) employed in MLD and ALD/MLD processes together with vapor pres- 


sures, P, at 100 °C and temperatures, T, corresponding to a vapor pressure of 2 mbar for some of the organic precursors (the values were calculated 




om the DIPPF 


! Project 801 database (full version) [58]). (continued) 






4,4'-oxydianiline 


ODA 


4,4'-oxydianiline; 4,4-diaminodiphenyl ether 






4-aminophenol 


AP 


4-aminophenol 






4-nitrobenzene-1 ,3-diamine 




2,4-diaminonitrobenzene 






7-octenyltrichlorosilane 


7-OTS 


7-octenyltrichlorosilane 






8-quinolinol 




8-hydroxyquinoline 


199 


100 


benzene-1 ,2,4,5-tetracarboxylic acid 




1 ,2,4,5-benzene tetracarboxylic acid; 
1 ,2,4,5-benzotetracarboxylic acid 




283 


benzene-1 ,2-dicarboxylic acid 




1 ,2-benzenedicarboxylic acid; 

1 9-hp-n7piHipnrhfiY\/lip apiH 




173 


benzene-1 ,3,5-tricarboxylic acid 




1 ,3,5-benzene tricarboxylic, 
1 ,3,5-benzotricarboxylic acid 






benzene-1, 3, 5-triol 




1 ,3,5-benzenetriol; phloroglucinol 






Del \Z.C\ lc- I ,0-UIOdl UUXyilO dOlU 




1 ,3-benzene dicarboxylic acid; 
1 ,3-benzodicarboxylic acid 




£.00 


benzene-1 ,4-dicarboxylic acid 




1 ,4-benzene dicarboxylic acid; 
1 ,4-benzodicarboxylic acid 




266 


benzene-1 ,4-diol 




1 ,4-benzendiol; hydroquinone 


9.39 


137 


benzoic acid 




benzoic acid 


1820 


71 


but-2-yne-1 ,4-diol 




2-butyne-1 ,4-diol 


•1 -1 Q 

1 To 


-i n7 


butane-1 ,4-diamine 




1 ,4-butane diamine 






butanedioic acid 




succinic acid 


0.569 


168 


decane-1,10-diamine 




1,10-diaminodecane 






decanedioic acid 




decanedioic acid; sebacic acid 


0.0466 


198 


decanedioyl dichloride 




sebacoyl dichloride 






ethane-1,2-diamine 


ED 


ethylenediamine 


56700 




ethane-1,2-diol 


EG 


ethylene glycol 


2100 


61 


ethanedihydrazide 




oxalic dihydrazide 






ethanedioic acid 




ethanedioic acid; oxalic acid 


50.9 


117 


ethanetetracarbonitrile 




tetracyanoethylene 






furan-2,5-dione 




maleic anhydride 


3260 


48 


fnmT^ 4-fir91hpn7nfi ir^n-1 *3 *-! 7-tp-trnnp 

I Ul U[0, t + I J L^JUCI 1Z.U IUIdM I,J,J,f iGll (J I lc 


PMDA 


pyromellitic dianhydride; 

1 ,2,3,5-benzenetetracarboxylic anhydride 






heptane-1 ,7-diamine 




1,7-diaminoheptane 






heptanedioic acid 




heptanedioic acid; pimelic acid 


0.806 


175 


hexa-2,4-diyne-1 ,6-diol 


HDD 


2,4-hexadiyne-1,6-diol; hexadiyne diol 






hexane-1,6-diamine 




1 ,6-hexanediamine; 1 ,6-diaminohexane; 


3670 


43 




hexamethylene diamine 


hexanedioyl dichloride 




adipyl dichloride; adipoyl chloride 






A/-(2-aminoethyl)ethane-1 ,2-diamine 




diethylenetriamine 


2610 


52 


A/,A/-bis(2-aminoethyl)ethane-1 ,2-diamine 




triethylenetetramine 


62.5 


117 


nonanedioyl dichloride 




azelaoyl dichloride 






octane-1,8-diamine 




1,8-diamino-octane 






octanedioic acid 




octanedioic acid; suberic acid 


0.109 


185 


octanedioyl dichloride 




suberoyl dichloride 






oxiran-2-ylmethanol 


GLY 


glycidol 


11400 


27 


pentanedioic acid 




pentanedioic acid; glutaric acid 


3.26 


160 


propane-1 ,2,3-tricarboxylic acid 




tricarballylic acid 






propane-1,2,3-triol 


GL 


glycerol 


25.6 


131 


propanedioic acid 




propanedioic acid; malonic acid 


4.55 


147 


propanedioyl dichloride 




malonyl chloride 






terephthalaldehyde 




terephthalaldehyde 


320 


94 


terephthalic acid bis(2-hydroxyethyl) ester 




terephthalic acid bis(2-hydroxyethyl) ester 


0.00091 


243 



1109 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



Table 1 : Organic compounds (and their different names and abbreviations) employed in MLD and ALD/MLD processes together with vapor pres- 
sures, P, at 100 °C and temperatures, T, corresponding to a vapor pressure of 2 mbar for some of the organic precursors (the values were calculated 

3 ro 



terephthaloyl dichloride 
tris(2-aminoethyl)amine 
tris(2-hydroxyethyl)amine 



terephthaloyl dichloride; terephthaloyl chloride 172 103 
tris(2-aminoethyl)amine 

triethanolamine 1.90 168 



pressures at 100 °C together with the temperatures corres- 
ponding to a vapor pressure of 2 mbar, which is a rather typical 
pressure used for ALD/MLD depositions. 

Organic thin films 

So far MLD has been used to produce polyamide, polyimide, 
polyimide-amide, polyurea, polyurethane, polythiurea, poly- 
ester and polyimine thin films. In Table 2, we list the character- 
istic polymer linkages seen in these films. For example, 
polyamides are polymers in which the precursors employed are 
combined with each other via amide bond formation whereas 
polyureas contain the urea linkage. And like the name of poly- 
imide-amides suggests, these polymers contain both an imide 
and an amide group. In the following sub-chapters all the 
different types of polymer thin films deposited by MLD until 
now are shortly presented. 

Polyamides 

Polyamides are extremely durable and strong, making them 
useful materials for a wide variety of applications, e.g., textiles, 
automotive industry applications and electronics. Polyamides 
can be classified to be aliphatic, semi-aromatic or aromatic, 
depending on the composition of the polymer chain. The 
polyamides produced by MLD have been deposited by using di- 
amines and acyl dichlorides as precursors (Table 3). The 
majority of these polyamides are aliphatic; so far only one semi- 
aromatic and one aromatic polyamide have been fabricated. 



Table 2: Characteristic linkages for the polymer types deposited using 
MLD. 



polymer 



polyamide 



polyimide 



polyurea 



polyurethane 



polythiourea 



polyester 



polyimine 



characteristic linkage 



R" 

O O 
R^N^R 



R' R'" 

r JV r ' 

R" 

%-V r " 

R R'" 
O 

R^R 



Aliphatic polyamides, i.e., nylons, have been deposited by 
MLD from precursors with a wide range of chain lengths. The 
shortest monomers employed are hexanedioyl dichloride and 
hexane-l,6-diamine which have been used to make nylon 66 
[8,50]. In the earlier work done by Shao et al. [8] the films were 
grown through polycondensation. The in situ FTIR studies by 
Du et al. [50] indicate that the deposition of nylon 66 displays 
characteristic ALD-type growth, i.e., linearity and self-limiting 
growth with the different precursor exposure lengths, although 
some ammonium chloride salt formation was observed. The 
highest GPC values for nylon 66 were 13.1 A per cycle when 
deposited at 60 °C on pretreated Si(100) [8] and up to 19 A per 
cycle on KBr substrates at the deposition temperature of 83 °C 
[50]. The latter value is somewhat higher than what the 
predicted unit-chain length, 17.4 A, would suggest: This higher 



than predicted growth rate was attributed to a CVD-type 
growth. Kubono et al. [6] deposited nylon 79 from heptane- 1,7- 
diamine and nonanedioyl dichloride at room temperature. The 
GPC value achieved, 18 A per cycle, was quite close to the 
calculated length of the repeating unit of the polyamide, i.e., 
22 A. Nagai et al. [7] fabricated several series of different 
nylons, including systems where up to four precursors were 
used, but the growth rates of the films were not discussed. 

Peng et al. [59] used butane- 1 ,4-diamine and terephthaloyl 
dichloride to grow semi-aromatic polyamide thin films. The 
highest achieved growth rate of this type of polyamide on 
Si(100) substrates was only 2 A per cycle at 85 °C. The near- 
edge X-ray absorption fine structure spectroscopy measure- 
ments showed that the oligomer units of the films were not 



1110 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 




precursor B 



references 



hexanedioyl dichloride 
O 

cr - ' - " Cl 

nonanedioyl dichloride 

Ck /\ /\ /\ XI 



hexane-1,6-diamine 



H 2 NL 



"NH 2 



heptane-1 ,7-diamine 
H 2 N N H 2 



[8,50] 



[6] 



hexanedioyl dichloride 
O 

CI 



CI 



octanedioyl dichloride 
O 




decanedioyl dichloride 




hexane-1,6-diamine 



H,N 



octane-1,8-diamine 



decane-1,10-diamine 



[7] 



terephthaloyl dichloride 
C> /=N CI 

terephthaloyl dichloride 
O /=\ CI 



butane-1,4-diamine 
NH 2 



H 2 N 

1 ,4-diaminobenzene 



[59] 



[60] 



perpendicular to the substrate surface but significantly tilted, 
suggesting that double reactions take place during the growth. 

The only aromatic polyamide grown by MLD so far was fabri- 
cated by using terephthaloyl dichloride and 1,4-diaminoben- 
zene as precursors. In industry, these two chemicals are used as 
precursors for a mechanically strong and thermally stable 
polymer known as Kevlar. The growth rate for the tere- 
phthaloyl dichloride+l,4-diaminobenzene MLD process was 
not constant: At 145 °C the measured GPC varied between 0.5 
and 3.3 A per cycle, which is considerably less than what could 
be assumed from the calculated chain length, which is 12.9 A. 
The low growth rate was assumed to be due to the non-ideal 
polymer-chain orientation: rather than standing up the polymer 
chains were suggested to have a more parallel orientation 
towards the substrate. [60] 



Polyimides 

As with the polyamides discussed above, polyimides can be 
classified to be aliphatic, semi-aromatic or aromatic depending 
on the chain composition. Furo[3,4-/][2]benzofuran- 1,3,5,7- 
tetrone (pyromellitic dianhydride, PMDA), widely used as a 
raw material for polyimides, has also been used as a precursor 
in all the MLD works on polyimide thin films so far (Table 4). 

Putkonen et al. [61] combined PMDA together with ethane-1,2- 
diamine (ED) and hexane-l,6-diamine to produce semi- 
aromatic polyimides. The highest growth rates achieved for 
these thin films were 3.9 A per cycle for the former and 5.8 A 
per cycle for the latter case, both at the deposition temperature 
of 160 °C. As the chain lengths for the ED and hexane-l,6-di- 
amine containing polyimide units are 9.8 and 14.5 A, respect- 
ively, the GPC values achieved are quite far from those 



1111 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 




precursor A precursor B precursor C references 



furo[3,4-f][2]benzofuran-1 ,3,5,7-tetrone ethane-1 ,2-diamine 



O O 




[61] 



O O 

furo[3,4-f][2]benzofuran-1 ,3,5,7-tetrone hexane-1 ,6-diamine 



O O 




[61,62] 



O O 

furo[3,4-/][2]benzofuran-1 ,3,5,7-tetrone 4-nitrobenzene-1 ,3-diamine 



O O NH 2 




[5] 



O O N0 2 

furo[3,4-/][2]benzofuran-1 ,3,5,7-tetrone 4,4'-oxydianiline 




[5,61,63-66] 



O O 

furo[3,4-/][2]benzofuran-1 ,3,5,7-tetrone 1 ,4-diaminobenzene 



O O 




[61,67] 



O O 

furan-2,5-dione 1 ,4-diaminobenzene furo[3,4-/][2]benzofuran-1 ,3,5,7-tetrone 

O O 

O^V^O H 2 NH^NH 2 

o o 




expected for an ideal MLD growth [61]. In a later study Salmi 
et al. [62] also grew polyimide films from PMDA and hexane- 
1,6-diamine. The depositions were done at the temperature of 
170 °C and a GPC of 5.6 A per cycle was reported. The article 
concentrates on nanolaminates fabricated from Ta20s and the 
polyimide and will be discussed in more detail later in this 
review [62]. 

In all the other polyimides deposited by MLD the second 
precursor used has been aromatic. When 4-nitrobenzene-l,3-di- 
amine was employed together with PMDA, no strong bonds 



formed between the two species, preventing the successful film 
growth [5]. 

The aromatic diamine precursor 4,4'-oxydianiline (ODA) has 
been used by several research groups. The highest growth rates 
reported were obtained by Yoshimura et al. [5]: In their experi- 
ments a GPC value as high as 6.7 A per cycle on an unspeci- 
fied substrate was achieved. In later experiments carried out by 
Putkonen et al. [61] the highest GPC value obtained was 4.9 A 
per cycle on Si(100) and soda lime glass substrates at 160 °C. 
Yoshidaet al. [63,64] employed Au-coated Si substrates, modi- 



1112 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



fied with 4-aminothiophenol to obtain NH2-group terminated 
surfaces. They conducted the experiments at 170 °C, achieving 
a GPC value of 2.0 A per cycle. As the approximate length of 
the PMDA-ODA chain is 14.9 A, none of the groups achieved 
GPC values close to full monolayer coverage. Yoshida et al. 
[63] also modified the films electrically by using a scanning 
probe microscope and reported significant increases in the 
conductivity of the films. The precursor combination, 
PMDA+ODA, was also employed by Haq et al. [65], who used 
reflection-absorption infrared spectroscopy to study the growth 
on Cu(l 10) surfaces as a function of temperature and coverage 
but did not discuss growth rates of the films. Also Miyamae et 
al. [66] deposited thin films by using PMDA and ODA, but as 
PMDA was introduced to the reactor so that the pulsing of 
ODA was kept on, the growth process was not necessarily of 
the real MLD type. 

The combination, PMDA+l,4-diaminobenzene, was used to 
fabricate organic thin films by Putkonen et al. [61], and also by 
Bitzer and Richardson [67]. The former group achieved the 
highest GPC value of 1 .4 A per cycle, which is well below the 
calculated chain length of 14 A, when using Si(100) and soda 
lime glass substrates at the deposition temperature of 160 °C. 
The latter group fabricated ultrathin films on Si(100)-2xl at 
room temperature but no GPC value was reported. Bitzer and 
Richardson [68] later also deposited ultrathin films on Si(100)- 
2x1 first functionalized with maleic anhydride, followed by 
stepwise exposures of 1 ,4-diaminobenzene and PMDA at room 
temperature. 

Polyimide-amides 

Miyamae et al. [66] deposited polyimide-amide thin films by 
using PMDA and terephthaloyl dichloride as the first and third 
building blocks, whereas ODA or decane-l,10-diamine were 



used as the second one (Table 5). However, as with the 
PMDA-ODA films grown by the same group, also with the 
polyimide-amides the precursor pulses were overlapping and 
thus the process was not precisely of the MLD type. 

Polyureas 

Polyureas are tough elastomers with a high melting point. They 
are especially useful as protective coatings. The polymers form 
with a reaction between an isocyanate and an amine. Depending 
on the diisocyanate used to fabricate the polymer, polyureas can 
be divided to either aromatic or aliphatic systems. The aromatic 
polyureas are typically sensitive to light, changing color after 
exposure, whereas the aliphatic polyureas retain their color 
when treated similarly. Both types have been deposited by using 
MLD (Table 6). 

1,4-diisocyanatobenzene (1,4-phenylene diisocyanate, PDIC) is 
the only aromatic diisocyanate so far used to fabricate polyureas 
by MLD. Kim et al. [69] employed ED as the second precursor, 
depositing the films on Ge(100)-2xl at room temperature. The 
growth of the ultrathin films was investigated with multiple- 
internal-reflection Fourier-transform infrared spectroscopy, 
demonstrating the formation of urea linkages, although some 
imperfections in the growth, possibly due to double reactions of 
ED, were also observed. In a later study by Loscutoff et al. [70] 
the same precursors were used to fabricate thin films on 
Si(100), treated with 3-aminopropyltriethoxysilane before depo- 
sitions to get amine-terminated surfaces. The deposition 
temperatures used ranged from 25 to 100 °C, and the GPC of 
the films decreased from 4.1 A per cycle at 25 °C to 0.4 A per 
cycle at 100 °C. A full monolayer coverage was not achieved as 
the highest GPC value was well below the PDIC+ED molecule 
length of 12.6 A, and also below the somewhat lower GPC 
value of 6.5 A per cycle expected based on the density func- 



Table 5: Precursors used to deposit polyimide-amide thin films. 


precursor A 


precursor B 


precursor C 


reference 


furo[3,4-/][2]benzofuran-1,3,5,7-tetrone 


4,4'-oxydianiline 


terephthaloyl dichloride 






O 




O /=\ CI 








H 2 N^^O^^NH 2 


[66] 


0 


0 








furo[3,4-f][2]benzofuran-1,3,5,7-tetrone 


decane-1 , 1 0-diamine 


terephthaloyl dichloride 






0 














O /=\ CI 


[66] 


O 


0 









1113 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



precursor A 



precursor B 



references 



1 ,4-diisocyanatobenzene 



0=C=NH^y)-N=C=0 
1 ,4-diisocyanatobenzene 



Q=C=N^ ^N=C=Q 

1 ,4-diisocyanatobutane 

O^C^N— -^ N=C = 0 
1 ,4-diisocyanatobutane 

0=C = N^— N=C = ° 
1 ,4-diisocyanatobutane 

0=C = N^-- N=C = 0 
1 ,4-diisocyanatobutane 



0=C=N' 



,N=C=0 



ethane-1,2-diamine 
H 2 N V 



NH P 



2,2'-(propane-2,2-diylbis(oxy))-diethanamine 

H 2 N^\ o y o ^^NH 2 

ethane-1,2-diamine 

H * N ^NH 2 
N-(2-aminoethyl)ethane-1,2-diamine 



H 2 N 



"HH 7 



A/,W-bis(2-aminoethyl)ethane-1,2-diamine 

,NH 2 



H 2 N 



H 



-N 
H 



tris(2-aminoethyl)amine 
H 2 N^^ N /^NH 2 

NH 2 



[69-71] 



[72] 



[73] 



[73] 



[73] 



[73] 



tional theory calculations carried out by the group. The film 
growth was observed to be linear. The films were stable in 
ambient air and when annealed at least up to 250 °C [70]. 
Recently, Prasittichai et al. [71] achieved a GPC of 5.3 A per 
cycle when growing PDIC+ED films at room temperature. In 
their study self-assembled monolayers fabricated by using 
octadecyltrichlorosilane were used to prevent the growth of the 
PDIC+ED polymer, enabling the growth of patterned 3D struc- 
tures [71]. 

Zhou et al. [72] used 2,2'-(propane-2,2-diylbis(oxy))- 
diethanamine together with PDIC, and fabricated the films on 
Si(100) at room temperature. The growth rate of these films was 
ca. 6.5 A per cycle, which is considerably less than the chain 
length of the expected molecule (18 A). The study demon- 
strated that when treated with acid, the backbone of the formed 
film reacted from the acid-labile groups. When exposed to basic 
solution, the polymer films were stable. These experiments 
proved that MLD can be utilized to fabricate photoresist ma- 
terials. To make the film a photoresist material, a treatment with 
triphenylsulfonium triflate, a photoacid generator, was required 
after the deposition. 



ethyl)ethane-l,2-diamine and tris(2-aminoethyl)amine have 
been used as the second precursor, yielding cross-linked 
polyurea films with GPCs of 6.3, 6.7, 3.2, and 3.1 A per cycle, 
respectively. The depositions were carried out at room tempera- 
ture and Si(100) was used as the substrate. As the calculated 
chain lengths for the respective systems are 13.5, 17.2, 20.9, 
and 17.2 A, the growth rates of the polymers were considerably 
less than a full monolayer per cycle [73]. 

Polyurethanes 

As with polyureas, the reaction in which polyurethane forma- 
tion takes place involves an isocyanate, but instead of an amine, 
an alcohol is required for the urethane linkage. In the MLD- 
grown polyurethanes the isocyanate has been PDIC, whereas 
but-2-yne-l,4-diol and terephthalic acid bis(2-hydroxyethyl) 
ester has been used as the second precursor in experiments done 
by Lee et al. [74] (Table 7). As the aim of the group was to 
make templates for the synthesis of zeolites, the growth process 
of the films is not discussed in detail. However, from the thick- 
ness evaluations for the system PDIC+but-2-yne-l,4-diol, 
which vary from 9.5 to 6.1 A per cycle, it can be seen that the 
growth is not linear. 



Also aliphatic polyureas have been fabricated by MLD using 
only one diisocyanate, namely 1,4-diisocyanatobutane. ED, 
7V-(2-aminoethyl)ethane-l,2-diamine, N,iV-bis(2-amino- 



Polythiourea 

The only MLD-grown polythiourea was fabricated by using 
1,4-diisothiocyanatobenzene and ED as the precursors 



1114 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 




precursor B 



reference 



1 ,4-diisocyanatobenzene 



o=c=nH^^n=c=o 

1 ,4-diisocyanatobenzene 



0=C=N^\ /^N=C=0 



but-2-yne-1,4-diol 
HO 



terephthalic acid bis(2-hydroxyethyl) ester 



[74] 




[74] 



cr 



(Table 8). The deposition temperature in these experiments was 
kept at 50 °C and the films were grown on silicon and silica 
nanoparticles. The obtained growth rates were 1.9 and 2.8 A per 
cycle, respectively. These values are modest when comparing to 
the calculated chain length of 13.8 A [75]. 

Polyesters 

Polyesters are commonly made with reactions between various 
acids and alcohols. The only MLD polyester reported is poly- 
ethylene terephthalate (PET), which is an industrially widely 
used polymer. The precursors employed for the fabrication of 
PET films were terephthaloyl dichloride, which is an acid chlo- 
ride, and ethane- 1,2-diol (ethylene glycol, EG), a diol (Table 9). 
The deposition temperature range investigated was 145-175 °C. 
The films were grown on Si(100) substrates, cleaned ultrasoni- 
cally in ethanol and water. The depositions were carried out on 
both cleaned substrates and on substrates that were further func- 
tionalized with amine groups using (3-aminopropyl)triethoxysi- 
lane. The growth rates were significantly better with the func- 
tionalized substrates. The highest growth rate, 3.3 A per cycle, 
was obtained at 145 °C. The length of a PET molecule is 1 1 A, 
so only partial layer coverage was achieved, though [76]. 

Polyimines 

Terephthalaldehyde and 1,4-diaminobenzene have been used to 
form polyimine thin films (Table 10) at room temperature 



Table 9: Precursors used to deposit polyester thin films. 


precursor A 


precursor B 


reference 


terephthaloyl dichloride 


ethane-1,2-diol 




C> /=\ CI 


HO^° H 


[76] 



[9,77,78]. The growth process was investigated by using 
Au-coated glass substrates with a self-assembled monolayer of 
11-amino-l-undecanethiol. According to the investigations by 
Yoshimura et al. [78] the polymer wires grew in upward direc- 
tions. Later experiments showed that that the first six cycles 
yielded a GPC value of 10 A per cycle, after which the growth 
rate started to decline [77]. Different combinations of 
terephthalaldehyde, 1,4-diaminobenzene and ethanedihydrazide 
have been used to fabricate quantum dots of varying lengths: 
These thin films showed promise for sensitization in photo- 
voltaic devices [79,80]. 

Hybrid inorganic-organic thin films 

Since the first publications featuring inorganic-organic hybrid 
thin films, the number of articles relating to this type of films 
has been rapidly increasing. In the following subchapters the 
inorganic-organic thin films deposited by using the combined 









precursor A 


precursor B 


reference 


1 ,4-diisothiocyanatobenzene 
S=C=N^^^N=C=S 


ethane-1,2-diamine 
H2N ^NH 2 


[75] 



1115 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



Table 10: Precursors used to deposit pol 
precursor A 



precursor B 



precursor C 



references 



terephthalaldehyde 

KK 

terephthalaldehyde 
O /=\ H 



1 ,4-diaminobenzene 



1,4-diaminobenzene 



ethanedihydrazide 
,, O 

kl ,NH 2 



[9,77,78] 



[79,80] 



ALD and MLD technique are presented. The different thin films 
produced are divided by the organic precursor employed. The 
subchapters "Alcohols and phenols", "Acids" and "Amines" 
contain all the different hybrid films made by using an organic 
precursor with hydro xyl, carboxyl or nitrogen atom function- 
ality, respectively. It should be noted that some of precursors 
under "Alcohols and phenols" have also another functional 
group, such as 4-aminophenol with both -OH and -NH2 
groups. (2S)-2-Aminopentanedioic acid, an amino acid with 
carboxylic acid functionality, is considered under "Acids". 
"Other organic precursors" consist of all the organic precursors 
that do not fall clearly into the categories mentioned above. The 
thin films based on 7-octenyltrichlorosilane are presented in 
their own subchapter at the end of this chapter. 

Alcohols and phenols 

Most of the inorganic-organic hybrid thin films reported so far 
have been deposited by using alcohols or phenols as the organic 



precursor (Table 11). When metal precursors are combined with 
alcohols or phenols, the resultant material is often called 
"metalcone" [27]. Most of the research has been focused on the 
aluminum-based alucones [12,13,15,16,38,39,81-97], but also 
zincones [13,33,38,39,46,51,52,56,94,95,98-103], titanicones 
[53,55,95,104,105], and zircones [54,87] have been studied. 
The schematic structure of ideally growing TMA+ethane- 1 ,2- 
diol, DEZ+benzene-l,4-diol and TMA+oxiran-2-ylmethanol 
hybrid films are shown in Figure 6 as examples of these types 
of films. 

Ethane- 1,2-diol (ethylene glycol, EG) is the simplest diol used 
to fabricate inorganic-organic hybrid materials by ALD/MLD. 
The use of EG together with TMA was first reported by 
Dameron et al. [12]. The thin films were grown in the tempera- 
ture range of 85-175 °C. The GPC value for the films varied 
from 4.0 to 0.4 A per cycle, decreasing with increasing deposi- 
tion temperature. The TMA+EG hybrids were unstable in 



cursors used to deposit inorganic-organic hybrid thin films, 
organic precursor inorganic precursor 



ethane-1,2-diol 
HO^-° H 



propane-1,2,3-triol 



AI(CH 3 ) 3 
Zn(C 2 H 5 )2 

T1CI4 
Zr(C 4 H 9 0) 4 



AI(CH 3 ) 3 
Zn(C 2 H 5 ) 2 
TiCI 4 



references 

[12,15,16,81-89,91,92,95,96] 
[51,52,98,101] 
[53,104,105] 
[54,87] 



[95] 
[95] 
[53,95,105] 



hexa-2,4-diyne-1 ,6-diol 
HO^ 

OH 



Zn(C 2 H 5 ) 2 
TiCI 4 



[46,56] 
[55] 



1116 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



11: Alcohol and phenol precursors used to deposit inorganic-organic hybrid thin films, (continued) 



benzene-1,4-diol 
.OH 



HO 



AI(CH 3 ) 3 
Zn(C 2 H 5 ) 2 



[13,94] 
[33,95,99-103] 



benzene-1,3,5-triol 
OH 



HO 




OH 



AI(CH 3 ) 3 



[13] 



oxiran-2-ylmethanol 
U^OH 



4-aminophenol 



H 2 N- 



OH 



AI(CH 3 ) 3 
Zn(C 2 H 5 ) 2 



Zn(C 2 H 5 ) 2 
TiCI 4 



[38,39,90,93,97] 
[38,39] 



[36,103,106] 
[37] 



8-quinolinol 




tris(2-hydroxyethyl)amine 
HO^ N ^OH 

s 



OH 



AI(CH 3 ) 3 
Zn(C 2 H 5 ) 2 

TiCU 



AI(CH 3 ) 3 



[13,35] 
[13,35] 

[13,35] 



[107] 



2-aminoethanol + furan-2,5-dione 
,NH 2 + 



HO 



O^ x O ^O 



AI(CH 3 ) 3 



[16,48,84,108,109] 



2-aminoethanol + propanedioyl dichloride 
O O 

^NH 2 + ju 



HO' 



cr 



-ci 



TiCI 4 



[49] 



ambient conditions, but capping with Si02 improved the 
stability. Mechanical studies of the TMA+EG hybrids showed 
that the material is brittle, with a toughness of about 
0.17 MPa-m 0 5 [81]. The brittleness was also observed with 
nanointendation measurements, giving an elastic modulus of 
about 37 GPa and a Berkovich hardness of about 0.47 GPa [84]. 
However, when the TMA+EG hybrid with an additional H2O 
pulse was employed as an interlayer between ALD-grown 
AI2O3 and a Teflon substrate, it was noticed that the stress 
caused by the difference in the coefficient of thermal expansion 



between the coating and the substrate was significantly reduced, 
preventing the cracking of the AI2O3 coating [96]. 

The first zincones were fabricated by combining EG with DEZ. 
This type of hybrid was reported by Peng et al. [51] and Yoon 
et al. [52]. The former group deposited the thin films in the 
temperature range of 100-170 °C. The maximum GPC value, 
about 0.57 A per cycle, was achieved for the film deposited at 
120 °C, while the lowest GPC, about 0.39 A per cycle, was 
observed for the films deposited at 165 °C. However, the 



1117 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 




Figure 6: Schematic illustration of ALD/MLD inorganic-organic hybrid thin films deposited by using (a) TMA with ethane-1 ,2-diol, (b) DEZ with 
benzene-1 ,4-diol, and (c) TMA with oxiran-2-ylmethanol. 



DEZ+EG hybrids are unstable in ambient air and the thickness 
measurements were carried out on reacted films with a reduced 
thickness. The films by Yoon et al. [52] were deposited at 
temperatures between 90 and 170 °C. Unlike with the deposi- 
tions by Peng et al. [51], the thickness of the films decreased 
with increasing temperature. A growth rate of 0.7 A per cycle 
was measured for films deposited at 130 °C. Both groups 
reported that although the DEZ+EG hybrid reacts with water in 
ambient air, the films remained stable after an initial quick reac- 
tion. Recently Liu et al. [101] grew DEZ+EG hybrid films at 
150 °C. They observed that the GPC varied strongly depending 
on the number of deposition cycles, from 0.39 to 0.86 A per 
cycle for 500 and 2000 cycles, respectively. This phenomenon 
was speculated to be due to the double reactions occurring 
during the growth. Also the GPC is far below the length of a 
Zn-EG unit, which was estimated to be around 6.9 A. Thermal 
conductivity of the DEZ+EG hybrid was measured to be around 
0.22-0.23 W/(m-K), with little variation in the values when 
thicker samples were analyzed. The volumetric heat capacity 
was about 2.5 J/(cm 3 -K) [101]. 

Ethylene glycol has also been used with TiCl4 to deposit titani- 
cone films [53,104,105]. Deposition temperatures for the 
TiCl 4 +EG films varied from 90 to 135 °C. The GPC first 
decreased gradually from 4.6 A per cycle at 90 °C to 4.1 A per 
cycle at 125 °C, after which there was a larger drop to 1 .5 A per 
cycle at 135 °C. The TiCLi+EG hybrid thin films were unstable: 
The thickness diminished by 15% over five days and after 
25 days the total reduction was 20%. The elastic modulus and 
hardness values measured by using nanointendation were 
extremely low, i.e., about 8 GPa and about 0.25 GPa, respect- 
ively [53]. 

Zircone thin films have been fabricated by using EG together 
with zirconium tert-butoxide (ZTB) [54,87]. The depositions 



were carried out in the temperature range of 105-195 °C. The 
GPC decreased with increasing deposition temperature, from 
1.6 to 0.3 A per cycle. The films were stable in ambient air: A 
decrease in thickness of only ca. 3% was observed after expo- 
sure in ambient air for one month [54]. 

Titanicone films have been also made by using propane- 1,2,3- 
triol (glycerol, GL) together with TiCl4. The deposition 
temperature regime used for these films was between 130 and 
210 °C. The GPC value at 130 °C was 2.8 A per cycle and 
dropped with increasing temperature gradually to 2.1 A per 
cycle at 210 °C. A small increase of film thickness was 
observed when thin films were exposed to air. The elastic 
modulus and hardness were both higher for the GL titanicones 
than those based on EG, with values of about 30.6 GPa and 
about 2.6 GPa, respectively. The films were also more ther- 
mally stable than the EG-based counterparts, probably due to 
crosslinking in the structure [53]. 

Hexa-2,4-diyne-l,6-diol (HDD) has been combined together 
with DEZ and TiCl4 to form cross-linked polydiacetylene struc- 
tures. These films were polymerized by using UV irradiation 
after the DEZ/TiC^ and HDD pulses. Deposition temperatures 
of 100-150 °C, and 100 °C were used for the Zn- and 
Ti-containing hybrid films, respectively. The calculated ideal 
layer thickness for both types of hybrids was approximately 
6 A. The GPC of the DEZ+HDD films was close to the ideal, 
i.e., 5.2 A per cycle, whereas the measured GPC for the 
TiCl4+HDD system reached full mono layer coverage, i.e., 6 A 
per cycle. Investigation of the electrical properties of the 
DEZ+HDD thin films revealed the films had an excellent field 
effect mobility (>1.3 cm 2 -V~'-s). When TEM was used to 
observe structures obtained with 50 cycles of TiC>2 and one 
cycle of TiCl4+HDD, individual nanolayers from each material 
could be seen [55,56]. 



1118 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



Benzene- 1,4-diol (hydroquinone, HQ), a phenol with two -OH 
groups in a para position, and TMA has been used to deposit 
films in the temperature range of 150-400 °C. The GPC was 
quite constant regardless of the deposition temperature used, 
around 3.5 A per cycle. When the films were stored in ambient 
air for one week, a decrease of 25% in film thickness was 
measured [13]. Zn-containing HQ films have also been fabri- 
cated with DEZ. The DEZ+HQ films were grown at tempera- 
tures between 130 and 170 °C, yielding GPC of 1.6 A per cycle 
at 150 °C. The stability of these films was not discussed [33]. In 
another study, a GPC of 2.7 A per cycle at the deposition 
temperature of 150 °C was obtained. Although considerably 
higher than when comparing to GPC obtained earlier, it is still 
far from the ideal growth, which was estimated to be around 
8.4 A per cycle. Thermal conductivity of the DEZ+HQ films 
was considerable higher than those of the DEZ+EG films, i.e., 
about 0.32-0.38 W/(m-K): The difference was attributed to a 
more vertical growth in case of the hybrid using the aromatic 
precursor when compared to the one using the linear precursor. 
Also the volumetric heat capacity was higher with the DEZ+HQ 
than with DEZ+EG, i.e., about 3.1 J/(cm 3 -K) [101]. 

Benzene- 1, 3, 5-triol, an aromatic compound with three -OH 
groups, has been used to fabricate hybrid thin films together 
with TMA. The depositions were carried out in temperature 
range of 175-400 °C. The GPC was about 5 A per cycle regard- 
less of the used deposition temperature. When stored in ambient 
air for one week, an increase of 9% in film thickness was 
measured [13]. 

There have been several studies featuring inorganic-organic 
hybrid materials deposited by using TMA and oxiran-2- 
ylmethanol (glycidol, GLY) [38,39,90,93,97]. The GLY mole- 
cule is heterobifunctional with both hydroxy and epoxy func- 
tionalities. Heterobifunctionality is believed to reduce the 
number of double reactions. During the deposition an epoxy 
ring-opening reaction takes place: The reaction is strongly 
dependent on the use of strong Lewis acids such as TMA. The 
first detailed studies of the growth of TMA+GLY thin films 
were published by Gong et al. [38] and Lee et al. [39] The 
temperature regimes used by the two groups were 90-150 °C 
and 100-175 °C, respectively. The GPC values reported varied 
greatly. Gong et al. [38] reported GPC values which decreased 
with increasing deposition temperature, the values varying from 
24 (which is considerably higher than the calculated chain 
length) to 6 A per cycle, whereas Lee et al. [39] reported a 
growth rate of only 1.3 A per cycle at 125 °C. Both groups 
reported that the TMA+GLY thin film growth is sensitive to 
reaction conditions. As the paper by Lee et al. [39] was the one 
published a little later, they speculated that the difference in 
their GPC values when compared to those reported by Gong et 



al. [38] could be due to CVD-type growth regime at lower 
deposition temperatures and with short purging times. Gong et 
al. [38] observed that the TMA+GLY films were relatively 
stable in ambient air: A 168 hour exposure resulted in the 
absorption of some OH~ according to FTIR data, but no change 
in thickness was detected. Both groups performed experiments 
also with DEZ+GLY systems, but no decent growth was 
observed for these films. It was concluded that DEZ as a weaker 
Lewis acid is not able to sufficiently catalyze the reaction 
required for the film growth to proceed [38,39]. 

The 4-aminophenol (AP) molecule is heterobifunctional 
consisting of a benzene ring with both -OH and -NH 2 groups. 
It has been investigated together with the inorganic precursor 
DEZ. The DEZ+AP hybrid thin film depositions were carried 
out at temperatures between 140 and 330 °C. Rather constant 
GPC of about 1.1 A per cycle was obtained at deposition 
temperatures of 140-200 °C, after which the GPC started to 
decrease with increasing temperature. The resultant films were 
stable when stored in ambient air when the humidity level was 
low [36]. Recently AP was used together with TiCl4, deposited 
at temperatures between 120 and 220 °C. The obtained GPC, 
which was 10-11 A per cycle in the deposition temperature 
range of 140-160 °C, was close to the value calculated from the 
bond lengths, 9.1 A per cycle. The films were relatively stable: 
Less than 10% decrease in film thickness was observed when 
stored in ambient air for 800 h [37]. 

Aromatic 8-quinolinol has been used together with TMA, DEZ 
and TiCLj [13,35]. The deposition temperature ranged from 85 
to 200 °C. The maximum growth rates were about 4, 6.5 and 
7.5 A per cycle for the A1-, Zn-, and Ti-containing hybrids, res- 
pectively, and were obtained at the lowest deposition tempera- 
ture used for each system. The growth rate gradually decreased 
with increasing temperature for all processes and no growth was 
observed in the films deposited at 200 °C. Significant photo lu- 
minescent activity was observed for the Al- and Zn-based 
hybrids whereas in the Ti-containing films a detectable but only 
small activity was perceived [35]. 

Tris(2-hydroxyethyl)amine, a tertiary amine, has been 
combined together with TMA to form hybrid thin films. The 
depositions were performed at 150 °C. The GPC was 1.3 A per 
cycle and the films were unstable in ambient air [107]. 

The use of more than two precursors can mitigate the proba- 
bility of double reactions. Yoon et al. [48] were the first to 
make inorganic-organic hybrid thin films by using three 
different precursors, namely TMA, 2-aminoethanol and furan- 
2,5-dione. From the two organic precursors, 2-aminoethanol is 
heterobifunctional while furan-2,5-dione is a ring-opening reac- 



1119 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



tant. The depositions were done in the temperature range of 
90-170 °C. The GPC was 24 A per cycle at 90 °C and 4.0 A per 
cycle at 170 °C. The films were not stable: It was observed that 
the films react with water mostly within the first 10 min of 
exposure to ambient air [48]. In later studies it was deduced that 
the growth process of the films is governed by the TMA diffu- 
sion into and out of the forming film, making the growth 
process strongly dependent on the TMA dose and purge times 
used. Also experiments carried out at 1 30 °C showed that the 
self-limiting growth could be only observed when the films 
remained thin [108]. According to nanointendation measure- 
ments conducted for the TMA+2-aminoethanol+furan-2,5-dione 
films deposited at 90 °C, the elastic modulus and Berkovich 
hardness were about 13 GPa and about 0.27 GPa, respectively 
[84]. 

Later, Chen et al. [49] also utilized three different precursors to 
fabricate hybrid thin films, i.e., TiCl/i, 2-aminoethanol and 
propanedioyl dichloride. Unlike with the TMA+2-amino- 
ethanol+furan-2,5-dione system, the growth sequence of these 
films did not consist of repeated introduction of the precursors 
in three stages, but in four: One full deposition cycle consisted 
of supplying TiCU, followed by 2-aminoethanol, then propane- 
dioyl dichloride and finally a repeated pulsing of 2-amino- 
ethanol. The heterobifunctionality of the 2-aminoethanol mole- 
cule was expected to improve reaction selectivity of the process. 



A GPC value of 6 A per cycle was achieved at deposition 
temperature of 100 °C when grown on carbon nanocoils. 

Acids 

Although the deposition of inorganic-organic hybrid thin films 
by using carboxylic acids as organic precursors was first 
mentioned by Nilsen et al. [13] already in 2008 (Table 12), the 
later articles published by the same group discuss the topic in 
greater detail [32,57,110], separately for saturated [57] and 
unsaturated [110] linear acids, as well as aromatic acids [32], 
when combined with TMA. The saturated linear acids were 
observed to form mostly bidentate complexes, and the unsatu- 
rated linear acids formed either bidentate or bridging 
complexes, whereas for the aromatic acids the complex type 
varied depending on the acid. Only one carboxylic acid has 
been used in combination with an inorganic precursor different 
from TMA, namely benzene- 1,4-dicarboxy lie acid with zinc 
acetate [111]. Representatives for the possible film structures 
are depicted in Figure 7. 

In the works reported for the combination, TMA+saturated 
linear acid, the chain length of the acid varies from two carbon 
atoms for the ethanedioic acid to ten carbon atoms for the 
decanedioic acid. Accordingly, the GPC values for the different 
systems also vary greatly (Figure 8), from 43 A per cycle 
achieved with the heptanedioic acid with seven carbon atoms to 



Table 12: Carboxylic acid precursors used to deposit inorganic-organic hybrid thin films 






organic precursor 


inorganic precursor 


references 


ethanedioic acid 






HO .0 


AI(CH 3 ) 3 


[13,57] 








propanedioic acid 






O O 
HO'^^OH 


AI(CH 3 ) 3 


[13,57] 


butanedioic acid 






HC V^OH 


AI(CH 3 ) 3 


[57] 


O 






pentanedioic acid 






O O 


AI(CH 3 ) 3 


[13,57,112] 








propane-1 ,2,3-tricarboxylic acid 






0 H °Y° OH 


AI(CH 3 ) 3 


[112] 


HO^^^^O 







1120 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



heptanedioic acid 
O O 



HO - v v v - 0H 
octanedioic acid 

O 

HO^ 




decanedioic acid 



HO 




(E)-butenedioic acid 
O 



"OH 



AI(CH 3 ) 3 



AI(CH 3 ) 3 



AI(CH 3 ) 3 



AI(CH 3 ) 3 



[13,57] 



[13,57] 



[13,57] 



[13,110] 



(Z)-butenedioic acid 
O O 



HO "OH 
(1E)-prop-1-ene-1,2,3-tricarboxylic acid 
O 



HO 
HO 



AI(CH 3 ) 3 



AI(CH 3 ) 3 



[13,110] 



[110] 



(2E,4E)-hexa-2,4-dienedioic acid 
O 

Ha - " " " ' OH 



AI(CH 3 ) 3 



[13,110] 



benzoic acid 
/ = \ OH 

benzene-1 ,2-dicarboxylic acid 
O 

OH 
OH 




AI(CH 3 ) 3 



AI(CH 3 ) 3 



[32] 



[13,32] 



benzene-1 ,3-dicarboxylic acid 
O O 



HO 




OH 



benzene-1 ,4-dicarboxylic acid 
HO /=\ 9 



AI(CH 3 ) 3 



AI(CH 3 ) 3 
Zn(CH 3 C0 2 ) 2 



[13,32] 



[13,32] 
[111] 



1121 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 



Table 12: Carboxylic acid precursors used to deposit inorganic-organic hybrid thin films, (continued) 



benzene-1 ,3,5-tricarboxylic acid 
O O 



HO 




OH 



O' °H 

benzene-1 ,2,4,5-tetracarboxcylic acid 
O O 



HO 
HO 




OH 
OH 



O O 
(2S)-2-aminopentanedioic acid 
O O 



HO 



OH 



NH, 



AI(CH 3 ) 3 



AI(CH 3 ) 3 



AI(CH 3 ) 3 



[13,32] 



[13,32] 



[112] 



(a) (b) (C) < 




Figure 7: Schematic illustration of ALD/MLD inorganic-organic hybrid 
thin films deposited using TMA together with (a) propanedioic acid 
(bidentate complex), (b) (E)-butenedioic acid (bridging complex), and 
(c) benzene-1 ,3,5-tricarboxylic acid (unbidentate complex). 

about 1 A per cycle for the decanedioic acid. From Figure 8, all 
the other systems but the one using heptanedioic acid and 
propane- 1, 2, 3-tricarboxylic acid show a decrease in GPC with 
increasing deposition temperature. For the heptanedioic acid the 
GPC is highly dependent on the deposition temperature, and the 
maximum value attained at 1 62 °C is even four times higher 
than the chain length of the acid precursor, a detail which could 
not be explained by the authors. The GPC of propane- 1,2,3-tri- 
carboxylic acid remained constant in the temperature range 
investigated. From Figure 8, the maximum GPC values for the 
acids containing 2-5 carbon atoms were of the same order of 
magnitude, whereas higher values were obtained for the acid 
systems with seven and eight carbon atoms and lower values for 
the decanedioic acid with ten carbons [57,112]. 



From Figure 8, when disregarding the heptanedioic acid+TMA 
hybrid, the GPCs for the unsaturated acids employed were 
higher when compared to those of saturated acids. The 
maximum GPCs varied between 7.8 and about 15 A per cycle 
obtained for (Z)-butenedioic acid and (l£)-prop-l-ene,l,2,3-tri- 
carboxylic acid systems. For all systems, the GPC decreased 
with increasing deposition temperature [110,112]. 

The aromatic carboxylic acids used to fabricate hybrid films 
together with TMA vary from each other regarding the number 
of carboxyl groups attached to the aromatic ring and how the 
groups are positioned. Benzoic acid has only one carboxyl 
group, which may explain why no decent growth for these films 
were observed. From dicarboxylic acids all the possible group 
positions were investigated, i.e., benzene- 1 ,2-dicarboxylic acid, 
benzene- 1,3-dicarboxy lie acid and benzene- 1,4-dicarboxy lie 
acid. The maximum GPC value for all these three acids, as seen 
from Figure 8, was in the range of 8 ± 1 A per cycle. The 
aromatic carboxylic acid containing three carboxylic groups, 
benzene-1, 3, 5-tricarboxylic acid, produced the highest GPC 
values from the aromatic acids: The maximum GPC was as high 
as 13.4 A per cycle. The maximum GPC of benzoic-1, 2,4,5- 
tetracarboxylic acid was considerably lower, 5.6 A per cycle. 
As can be seen from Figure 8, for all aromatic acid 
systems the GPC decreases with increasing deposition tempera- 
ture [32]. 

All the carboxylic acid+TMA hybrid films were stable: No 
changes were observed for the saturated and unsaturated acid 
systems when stored in ambient air for one year, whereas the 



1122 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 



20 



15 



saturated 
unsaturated 



10 - 



A. 
A 



100 



100 



o o 



o 



♦ 



□ ♦ 

X 



o 



A t 



o 



f 



150 



200 250 300 

Deposition temperature (C" 



150 



200 250 300 

Deposition temperature (C°) 



Oethanedioic acid (2C) 


50 r 


Apropanedioic acid (3C) 


45 - 


• butanedioic acid (4C) 


40 


■ pentanedioic acid (5C) 


35 - 


♦ propane-1 ,2,3-tricarboxylic 

acid (6C) 
□ octanedioic acid (8C) 


30 

25 ■ 


xdecanedioic acid (10C) 


20 


0(E)-butenedioic acid (4C) 


15 


A(Z)-butenedioic acid (4C) 


10 - 


■ (1 E)-prop-1-ene- 

1 ,2,3,tricarboxylic acid (6C) 


5 
0 



Aheptanedioic acid (7C) 




350 400 • (2E,4E)-hexa-2,4-dienedioic 

acid (6C) 



X benzoic acid 

■ benzene-1,2-dicarboxylic acid 

♦ benzene-1,3-dicarboxylic acid 
Abenzene-1,4-dicarboxylic acid 

• benzene-1,3.5-tricarboxylic acid 
Obenzene-1,2,4,5-tetracarboxylic acid 



350 400 



Figure 8: Growth per cycle values for inorganic-organic hybrid films deposited by using TMA with different carboxylic acids [32,57,1 10,1 12]. 



stability of aromatic carboxylic acids was observed for one 
week [32,57,110,112]. 

Salmi et al. [11 1] used benzene- 1,4-dicarboxy lie acid together 
with zinc acetate to deposit metal-organic framework (MOF) 
thin films. The films were deposited at 225-350 °C, and the 
GPC decreased with increasing deposition temperature, the 
highest value being 6.5 A per cycle. No growth was observed 
when the deposition was carried out at 350 °C. The as-deposited 
films were amorphous, but films deposited below 300 °C 
crystallized when kept at 60% humidity at room temperature. 
According to the time-of-flight elastic recoil detection analysis 
and FTIR experiments conducted for the crystallized films, the 
composition was close to that of MOF-5. 

Klepper et al. [112] used (25)-2-aminopentanedioic acid, an 
amino acid with a carboxylic acid functional group, together 
with TMA. The films were grown in the temperature range of 
200-350 °C. The GPC values decreased with increasing deposi- 



tion temperature. The highest GPC obtained was 20 A per 
cycle, which is considerably higher than the estimated chain 
length, which was 10 A. However, according to the XRD 
measurements, there was a peak around 4.5° in the 9-29 pattern, 
which corresponds to a d value of 19.9 A, indicating an Al-Al 
sheet distance, which is close to the maximum GPC. The group 
concluded that a sheet-like structure with an interplanary sheet 
distance of 20 A forms during the deposition of the film. Also 
these films were stable: No changes were observed visually or 
in XRR and FTIR measurements. 

Amines 

Two diamines, 1,4-diaminobenzene and ODA, have been used 
to fabricate inorganic-organic hybrid thin films by ALD/MLD 
(Table 13). The 1,4-diaminobenzene, i.e., an amine equivalent 
of HQ, was used together with TMA. The depositions were 
carried out in the temperature range of 200-400 °C. The GPC 
decreased with increasing deposition temperature, varying 
between about 1 and about 2 A per cycle. The obtained films 



1123 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 





is. 




organic precursor 


inorganic precursor 


references 


1 ,4-diaminobenzene 

4,4'-oxydianiline 
H 2 N 0 "^J^" N H2 


AI(CH 3 ) 3 
TiCI 4 


[34] 
[14,103,113] 



were unstable: An increase of ca. 30% in film thickness was 
observed when the films were kept in ambient air for two weeks 
[34]. 



of 14 s. The GPC of these films also increased with increasing 
temperature, from 0.8 A per cycle at 160 °C to 1.6 A per cycle 
at 310 °C [113]. 



The amine precursor ODA has been used together with TiCLj to 
deposit thin films at deposition temperatures varying from 160 
to 490 °C. GPC of these films increased with increasing deposi- 
tion temperature, from 0.3 A per cycle to 1.1 A per cycle with 
an ODA-pulse length of 3 s. The TiCl4 films were stable in 
ambient air when deposited at 250 °C and above. Wet-etching 
testing revealed that when treated with toluene, acetone, 
methanol, 1 M acetic acid, or water, no significant change in 
film thickness was observed [14]. However, as the ODA-pulse 
length of 3 s was not enough for the fully-saturated growth, in a 
later study TiCl4+ODA films with longer ODA pulse lengths 
were deposited. It was observed that an ODA pulse length of 
12-16 s led to saturated growth. Depositions at temperatures 
from 160 to 310 °C were carried out with an ODA pulse length 



Other organic precursors 

Only three organic precursors have been used that cannot be 
included in the alcohol/phenol, acid or amine categories, 
namely ethanetetracarbonitrile, which is, like the name indi- 
cates, a nitrile, 2-oxepanone, a cyclic ester, and 1,2- 
bis[(diamethylamino)dimethylsilyl]ethane, which is a hybrid 
material in itself (Table 14). 

Ethanetetracarbonitrile has been used together with vanadium 
hexacarbonyl V(CO)6 to produce the only vanadium-containing 
hybrid thin film deposited until now. The depositions were 
carried out at room temperature. A GPC of about 9.8 A per 
cycle was achieved. The film was a room-temperature magnet, 
with a large coercive field (ca. 80 Oe at 5 and 300 K). 



organic precursor 



ethanetetracarbonitrile 




inorganic/counter precursor 



reference 



1,2-bis[(dimethylamino)dimethylsilyl]ethane 



V(CO) 6 



AI(CH 3 ) 3 



0 3 



[114] 



[47] 



[115] 



1124 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



According to the experiments, the hybrid has a high T c and/or 
high thermal stability when compared to corresponding CVD- 
made films. It was concluded that the material holds promise 
for next-generation spin-related applications [114]. 

Gong et al. [47] used 2-oxepanone and TMA to make hybrid 
thin films in which the TMA as a Lewis acid catalyzes a ring- 
opening reaction. The films were deposited at 60-120 °C. The 
GPC value decreased with increasing temperature being 0.75 A 
per cycle at 60 °C GPC but only 0.08 A per cycle at 120 °C. 
The films were stable; no change in film thickness was 
observed when stored in ambient air for 30 days. 

The hybrid film fabricated by Zhou and Bent [115] differs in 
some aspects from the other hybrids made up till now. It is the 
only carbosiloxane film made so far and the precursor used is 
an inorganic-organic hybrid material in itself, namely 1,2- 
bis[(dimethylamino)dimethylsilyl]ethane. Both O3 and H2O 
were tried as the second precursor, but H 2 0 did not work as 
well as O3. It was speculated that double reactions hinder the 
growth when using H2O, whereas when O3 was used new 
surface groups were generated, improving the overall growth 
process. GPC of 0.2 A per cycle was achieved at the deposition 
temperature of 110 °C for the l,2-bis[(dimethylamino)di- 
methylsiryl]ethane+C>3 system. The films were extremely stable, 



withstanding well wet etching treatments of tetramethyl- 
ammonium hydroxide, HC1 and acetone as well as vacuum 
annealing: When annealed at 300 °C, no change in thickness 
was observed, 400 °C resulted in a thickness change of 6%, and 
after annealing at 600 °C, a thickness loss of 13 % was 
measured. 

7-octenyltrichlorosilane based thin films 

ALD/MLD has been used to deposit several different inor- 
ganic-organic hybrid films featuring 7-octenyltrichlorosilane 
(7-OTS) as one of the precursors (Table 15). The first article 
where 7-OTS was employed was published already in 2007 
[10]. However, 7-OTS in itself cannot be said to be a fully 
organic material as it contains silicon. Also 7-OTS was pulsed 
to the reactor together with water as a catalyst, not alone as in a 
conventional ALD/MLD process. The key stages for all these 
7-OTS hybrids during one cycle were the same: First 7-OTS 
and water were introduced, followed by an ozone treatment. 
The third stage consisted of metal precursor connection to the 
forming chain, followed by reaction with water. Schematic 
presentation of the forming chain is shown in Figure 9. The 
metal precursors used to make these type of hybrid thin films 
include Ti(OCH(CH 3 ) 2 ) 4 [10,40], TMA [46], Zr(C 4 H 9 0) 4 [43] 
and DEZ [44]. For all but TMA thin films the GPC was around 
10-11 A per cycle. The maximum growth rate for the 



Table 15: Inorganic-organic hybrid thin film systems based on 7-octenyltrichlor< 

precursors of organic layer precursors of inorganic layer(s) references 



7-octenyltrichlorosilane 
CI 

Cl-Si-CI 



H 2 0 



O3 Ti(OCH(CH 3 ) 2 ) 4 H 2 Q 



10,40 



7-octenyltrichlorosilane 
CI 

Cl-Si-CI 



H 2 0 



O3 



Ti(OCH(CH 3 ) 2 ) 4 
AI(CH 3 ) 3 



H 2 Q 



41,42 



7-octenyltrichlorosilane 
CI 

Cl-Si-CI 



H 2 0 



o 3 



Zr(C 4 H 9 0) 4 



H 2 Q 



43 



7-octenyltrichlorosilane 
CI 

Cl-Si-CI 



H 2 0 



o 3 



Zn(C 2 H 5 ) 2 



H 2 Q 



44 



7-octenyltrichlorosilane 
CI 

Cl-Si-CI 



H 2 0 



0 3 



AI(CH 3 ) 3 



H 2 Q 



45,46 



1125 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 



7-OTS+TMA film was only 6 A per cycle. There are some 
contradictions regarding the reported growth rates for a full 
monolayer covering. According to the earlier articles an ideal 
monolayer would have a length of 12 A [10,41] whereas a more 
recent one states that the full monolayer coverage would lead to 
a GPC value of 25 A per cycle [40]. 




Figure 9: Schematic illustration of an ALD/MLD inorganic-organic 
hybrid thin film deposited using 7-octenyltrichlorosilane and 
Ti(OCH(CH 3 ) 2 ) 4 . 

The 7-OTS+Ti(OCH(CH 3 ) 2 ) 4 containing films have been 
grown in a temperature range of 150-200 °C. When annealed, 
these films were stable up to 450 °C. When the electrical prop- 
erties of the films were investigated, the experiments showed 
that the current density decreased with increased film thickness. 
When the films were less than 6.7 nm thick, direct tunneling 
was observed while the 99 nm thick films had a good gate 
leakage resistance. The dielectric constant for the films was 
around 17 at 1 MHz. It was concluded that the system 



7-OTS+Ti(OCH(CH 3 ) 2 ) 4 is a good candidate for high-quality 
gate-insulating films on flexible substrates [10]. In later experi- 
ments rapid water permeation speed was observed for this type 
of film [40]. 

Although the first article mentioning the use of 7-OTS together 
with TMA was published already in 2008 [41], a more thor- 
ough investigation of the growth process was carried out only 
recently [45]. The GPC decreased with increasing temperature, 
from about 6 A per cycle at 100 °C to 2.5 A per cycle at 200 °C 
[45]. 

The Zr- and Zn-containing films were grown at 170 and 150 °C, 
respectively. These studies focus more on the nanolaminates 
fabricated with oxides and will be discussed in more detail later 
[43,44]. 

Processes with no growth details 

In this chapter ALD/MLD works about only thin films that 
provide no precise details regarding the experiments are shortly 
summarized. It should be noted that the organic precursors used 
in the growth processes are not included in any of the tables. 

The only Fe-containing hybrid films reported so far were grown 
by Smirnov et al. [11,116] from FeCl3 and 2-propyn-l-ol 
(propargyl alcohol, the simplest alcohol with an alkyne group). 
The deposition temperature was 200 °C, but no other experi- 
mental details were given. The focus was on the magnetic prop- 
erties of the films: Uncompensated antiferromagnetism was 
observed. 

Nilsen et al. [13] mention the use of nona-l,9-diol (1,9-nonane- 
diol) together with TMA at 200 °C, but besides a QCM figure 
there is no further information regarding the system. In the 
same article TiCl4 and ED were used to fabricate hybrid thin 
films. Only a QCM graph acquired at deposition temperature of 
75 °C and a cautioning to use long pulse lengths were given. 
They also combined 4-aminobenzoic acid with TiCl4 hoping 
that the carboxylic group would enhance the air stability of the 
hybrid film. No information regarding the depositions of these 
films was given but a QCM graph obtained at the deposition 
temperature of 200 °C. 

The deposition strategy consisting of four stages by Chen et al. 
[49] has been also used to fabricate Al-containing hybrid films 
from TMA. The organic precursors were the same as with the 
Ti-containing counterparts, i.e., aminoethanol and propanedioyl 
dichloride. 

Abdulagatov et al. [105] conducted pyro lysis studies on several 
hybrid films grown by using EG, GL, HQ and 2,3,5,6-tetrafluo- 



1126 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



robenzene-l,4-diol (tetrafluorohydroquinone) as the organic 
precursors: Al+GL, Al+EG, Al+HQ, Al+2,3,5,6-tetrafluoroben- 
zene-l,4-diol, Zn+GL, Zn+HQ, Zr+EG, Hf+EG and Mn+EG 
were deposited at 150 °C [105]. Preliminary FTIR studies 
measuring 2,2-dimethoxy-l,6-diaza-2-silacyclo- 
oxetane+ethylene carbonate and TMA+3-buten-l-ol containing 
systems have been shortly mentioned [60]. 

In his review article, George [26] mentions polydimethyl- 
siloxane films deposited by using H2O together with bis(di- 
methylamino)dimethylsilane, 1 ,3-dichlorotetramethyldi- 
siloxane and dimethylmethoxychlorosilane. The GPC of the 
films was negligible after 15 deposition cycles, so depositions 
consisting of four pulsing steps, i.e., TMA+H 2 0+dimethyl- 
methoxychlorosilane+H20, at 200 °C were tried as well. 
Although this process yielded linear growth with GPC of 0.9 A 
per cycle, the silicon content was too low. Hybrid films made 
by using TMA and l,4-diazabicyclo[2.2.2]octane (triethylenedi- 
amine) are also shortly mentioned. 

Tetrakis(dimethylamido)hafnium and EG have been combined 
to form hafnicones. Linear growth was observed at the deposi- 
tion temperature of 145 °C. GPC decreased from 1.2 A per 
cycle at 105 °C to 0.4 A per cycle at 205 °C. The films were 
said to be very stable [28]. 

Layer-engineered and nanostructured ALD/ 
MLD films 

Making various nanostructure and multilayer architectures by 
ALD and MLD is inherently easy due to the self-limiting cyclic 
growth process, which allows excellent control over the layer 
thickness and enables the conformal growth of the films. The 
ALD/MLD techniques have been successfully utilized to coat 
various nano structures, such as nanoparticles, nanowires and 
carbon nanotubes. Both post-deposition annealing and wet- 
etching processes have been employed to remove the organic 
part of the inorganic-organic hybrid material, leaving porous 
oxide backbones. By adding a second material to the deposition 
process, it is also possible to make thin-film mixtures, superlat- 
tices and nanolaminates with tailored properties. In the 
following subchapters depositions of such nanostructured and 
layer-engineered ALD/MLD materials are discussed. 

Fabrication of porous and nanostructured materials 

Hybrid ALD/MLD processes based on TMA and EG precur- 
sors have been used to coat nanoparticles and nanowires: TiC>2 
nanoparticles have been coated for photoactivity passivation 
[85], silica nanoparticles to fabricate AI2O3 films with 
controlled pore sizes (Figure 10a) [15,82], and CuO nanowires 
to obtain hollow A1 2 C>3 nanostructures [83]. Hybrid TMA+EG 
coatings have also been utilized to fabricate A^CVzeolite 



composite membranes. In the latter example the organic 
constituent was removed by means of a post-deposition oxi- 
dation to form microporous AI2O3 on and between zeolite crys- 
tals. The membranes showed promise for H2 separation [89]. 

Liang et al. [98] deposited DEZ+EG coatings on titania 
nanoparticles. The DEZ+EG hybrid was observed to reduce the 
photoactivity of the nanoparticles. When the films were 
annealed in air, porous ZnO was formed. However, the number 
of pores was rather low. Moreover, the size of the pores fluctu- 
ated widely. 

Porous TiC>2 films have been obtained both by post-deposition 
annealing and by treating TiCl4+EG films with UV light 
[53,104]. After the post-deposition annealing the films were 
observed to have both amorphous and crystalline states and ex- 
hibit high photocatalytic activity. In a more recent paper by 
Abdulagatov et al. [105], GL was used together with TiCl4 to 
deposit thin films which then were annealed to form porous 
structures. For the annealed films a significantly reduced sheet 
resistance was achieved, being as low as 2.2 * 10 2 Q. for films 
annealed at 800 °C; resistivity of the same films was 0.19 £l-cm. 
These films are anticipated to have some potential in applica- 
tions related to electrochemical reactions and electrochemical 
energy storage [105]. 

Gong et al. [38] deposited hybrid TMA+GLY coatings on 
electro-spun polyvinyl alcohol fiber mats, and subsequently 
annealed the samples at 400 °C for 48 h to remove the organic 
part. Conformality was verified by transmission electron 
microscopy, see Figure 10b. The pores formed were found to be 
mostly 5 A micropores, with some mesopores [38]. Lee et al. 
[39] investigated the effect of the annealing temperature for 
similar types of TMA+GLY films and found that annealing at 
350-500 °C removed the organic part of the films leaving 
A1 2 0 3 . 

Hybrid TMA+GLY thin films have also been used together 
with ALD-grown ZnO and TiC>2 to coat and protect electro- 
spun polyamide-6 (PA-6) nanofibers [90,93]. When TiC>2 was 
used alone, the process caused fiber degradation. It was 
revealed that whereas a layer of ZnO did not offer sufficient 
protection from the following TMA+H2O treatment, when a 
layer TMA+GLY was deposited after ZnO, it prevented further 
degradation of the fibers when a Ti02 coating was applied on 
the top [90]. 

Recently, Brown et al. [95] coated carbon nanotubes with 
several different metalcone materials, including the TMA+EG, 
TMA+GL, TiCl 4 +GL, and DEZ+GL systems, see Figure 10c. 
The measurement of mechanical properties revealed that the 



1127 




Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 




Figure 10: Electron microscope images of (a) 250 nm silica particles coated with a 25 nm thick layer of TMA+EG (reprinted with permission from [15], 
Copyright (2009) The Royal Society of Chemistry), (b) TMA+GLY film deposited on electro-spun polyvinyl alcohol fibers and annealed at 400 °C for 
48 h (reprinted with permission from [38], Copyright (201 1) American Chemical Society), and (c) metalcone coated carbon nanotubes (reprinted with 
permission from [95], Copyright (2013) American Chemical Society). 



MLD coatings caused a significant reduction in failure strain, a 
modest improvement in ultimate tensile strength, and a signifi- 
cant improvement in Young's modulus. The greatest failure 
strain, about 3.9%, was achieved for TMA+GL coated 
nanotubes. The highest average Young's modulus and the ulti- 
mate tensile strength were found for TMA+EG coated samples, 
with values of about 7 GPa and about 100 MPa, respectively. 
Young's modulus for the uncoated carbon nanotube was 
510 MPa, 2.2 GPa for 10 nm thick TMA+GL coated samples, 
and about 9 GPa for a composite coating consisting of 5 nm 
TMA+GL and 5 nm A1 2 0 3 . 

Hybrid TMA+GLY films have also been grown on hydrophobic 
polydimethylsiloxane (PDMS) silicone exposed to sequential 
vapor infiltration of TMA+water. The aim of the study was to 
produce PDMS with a hydrophilic surface. When pure AI2O3 
was used to coat the PDMS, a hydrophilic surface was obtained. 
However, after storing for 48 hours in ambient air the coating 
became more hydrophobic, losing the desired wetting character- 
istics. The TMA+GLY coating produced also a hydrophilic 
surface, which retained the hydrophilicity for more than two 
weeks in ambient air. However, the most hydrophilic and stable 
coatings were obtained by fabricating structures consisting of 
100 cycles of A1 2 0 3 and TMA+GLY [97]. 



Liang et al. [109] deposited TM A+tris(2-hydroxy- 
ethyl)amine+furan-2,5-dione thin films at 150 °C on 500 nm 
spherical silica particles. Carbon was removed from the films 
by soaking them in water for one day or by a 1 h oxidation in 
air at 400 °C. The resultant nanopores were 0.6-0.8 nm (and 
some 17 nm pores) and 0.8 nm in size, respectively. 

Dry-etching with oxygen of TMA+EG and TMA+tris(2- 
hydroxyethyl)amine+furan-2,5-dione hybrids and wet-etching 
of TMA+tris(2-hydroxyethyl)amine+furan-2,5-dione with 
various solvents has also been studied. It was observed that HC1 
could be used to etch the thin film in a controlled manner, 
which makes of TMA+tris(2-hydroxyethyl)amine+furan-2,5- 
dione hybrid films promising materials for MEMS/NEMS 
applications [16]. 

Films from TiCl4, tris(2-hydroxyethyl)amine and propanedioyl 
dichloride have been fabricated on suspended CuO nanowires 
and carbon nanocoils by using a four-deposition-stage ap- 
proach. The films were annealed at 600 °C in 5% H 2 /Ar for 2 h 
after the deposition. Excellent photocatalytic activity was 
observed for the nanoporous TiC^/carbon nanocoil structures. 
Nanoporous AI2O3 structures were also obtained using TMA as 
the inorganic precursor [49]. 



1128 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



Fabrication of thin-film mixtures, superlattices and 
nanolaminates 

The ALD/MLD thin-film mixtures, superlattices and nanolami- 
nates are all made by using at least two different materials, for 
example a purely organic material and an oxide, by varying the 
number of deposition cycles of each material. The distinction 
between the various types of such layer-engineered materials is 
not completely unambiguous. In principle, a mixture is formed 
when the number of deposition cycles of each material is kept 
relatively low; then, if no full monolayer coverage is achieved 
with each material, the materials do not form separate layers but 
instead a homogenous mixture. On the other hand, if the growth 
happens in a well-controlled manner, repeated cycles may lead 
to a superlattice structure. The nanolaminates are formed by 
using larger number of deposition cycles, so that at least one of 
the materials achieves nanometer scale thickness and the ma- 
terials form individual layers in the structure. 

Only three articles published so far feature multilayer struc- 
tures that contain purely organic nanoscale layers [62,70,72], 
Salmi et al. [62] fabricated Ta205/polyimide nanolaminate 
structures consisting of nanoscale inorganic Ta 2 Os and organic 
polyimide layers. Tantalum ethoxide and water were used to 
deposit Ta 2 05, and PMDA and hexane-l,6-diamine for the 
polyimide deposition. All nanolaminates were constructed from 
five bilayers, with three different constructions: 10 nm Ta 2 Os + 
10 nm polyimide (shown in Figure 11), 15 nm Ta 2 Os + 5 nm 
polyimide, and 5 nm Ta 2 Os + 15 nm polyimide. The study 
focused on dielectric and mechanical properties of these 
nanolaminates. It was observed that the permittivity of the 
nanolaminates increased with the Ta 2 Os content, but not in a 
linear manner. The trend was similar for the refractive index, 
but in a more linear way. It was also seen that the insulating 



Ta 2 O s Polyimide 




Substrate 



Figure 11: Field emission scanning electron microscopy image of a 
nanolaminate fabricated using five bilayers of 10 nm Ta20s and 10 nm 
polyimide (reprinted with permission from [62], Copyright (2009) 
WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim). 



properties of the parent materials could be improved through 
the fabrication of nanolaminate structures. Nanointendation 
measurements revealed that the softness, elasticity and plas- 
ticity of the films increased with increasing polyimide content. 

Loscutoff et al. [70] deposited purely organic nanolaminates at 
room temperature consisting of 30-50 nm thick PDIC+ED and 
PDIC+2-[(2-aminoethyl)sulfanyl]ethanamine layers. The 
growth of the latter material was not discussed in detail in the 
article, though. Depth profiles by a sputter technique were 
investigated on two different nanolaminates: The first consisted 
of two layers of 40 nm PDIC+ED, with a 50 nm thick PDIC+2- 
[(2-aminoethyl)sulfanyl]ethanamine layer in between, the 
second of two 30 nm thick PDIC+2-[(2-aminoethyl)sulfan- 
yl]ethanamine layers with a 40 nm thick PDIC+ED layer 
between them. Although a small unexpected sulfur signal indi- 
cated the presence of some unreacted sites, it could be 
concluded that the composition of the nanolaminates agreed 
well with the expected values. 

As part of their study on potential photoresist materials, Zhou et 
al. [72] also investigated nanolaminates of purely organic 
constituents, i.e., PDIC+ED and PDIC+2,2-(propane-2,2- 
diylbis(oxy))diethanamine. Organic films of the former type are 
stable in HC1, whereas of the latter type are not. Three-layer 
structures were deposited to investigate the stability of the 
nanolaminates in HC1: First 3, 6 or 9 cycles of PDIC+ED were 
deposited, followed by 3 or 6 cycles of PDIC+2,2-(propane-2,2- 
diylbis(oxy))diethanamine, finished with 3 cycles of PDIC+ED 
on the top. According to the thickness measurements, only the 
bottom layer remained after the acid treatment, indicating that 
the cleaving occurred at the positions of acid-labile groups. 
Also nanolaminates consisting of 3 cycles of PDIC+2,2- 
(propane-2,2-diylbis(oxy))diethanamine and 8 or 15 cycles of 
PDIC+ED on top were deposited. These films were soaked with 
triphenylsulfonium triflate, UV radiated, baked and developed. 
No significant changes in thickness were observed and it was 
concluded that the top layer behaves as a photoacid generator 
barrier. 

Quite many inorganic-organic hybrid compositions have been 
utilized to fabricate homogeneous thin-film mixtures, superlat- 
tice structures and nanolaminates. Lee et al. [86,87] fabricated 
mixtures consisting of TMA+EG alucone and Al 2 03 at 135 °C, 
and varied the oxide:hybrid cycle ratio from 1:3 to 6:1 to accu- 
rately control the density, refractive index, elastic modulus and 
hardness of the films. The values obtained varied between those 
for pure alucone and pure Al 2 03, for density from 1.6 to 
3.0 g/cm 3 , for refractive index from 1.45 to 1.64, for elastic 
modulus from 20 to 200 GPa, and for hardness from 1 to 
13 GPa. 



1129 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



The stability and optical properties of mixtures of alucone (from 
TMA+EG) and AI2O3 have been also studied and compared to 
those of pure alucone films. The depositions were carried out on 
Si(100) substrates at 150 °C. It was observed that when the 
sample was kept in air the thickness and refractive index of pure 
alucone decreased by about 20% and about 1 .4%, respectively, 
over the first three days, after which there were no further 
changes. When the alucone films were kept in a desiccator, 
there were no significant changes in thickness or refractive 
index. It was also revealed that capping a 100 nm thick alucone 
layer with a 20 nm thick AI2O3 layer improved the stability, as 
did the fabrication of a nanolaminate structure consisting of five 
bilayers of 20 nm of alucone and 10 nm of AI2O3. When homo- 
geneous alucone:Al 2 C>3 mixtures were investigated, the mix- 
ture with a ratio of 5:1 lost 12-17% of its initial thickness in 
open air, whereas the 5:5 mixture was stable, and the thickness 
of the 1:1 mixture was reduced by 4%. The change in refractive 
index for the 5:1 mixture was larger than for the pure alucone 
film. Measurements performed on annealed films showed that 
the refractive index could be tuned by the speed of the 
annealing: Slow annealing resulted in a film with more pores 
and less collapse in the film. When heated, the 1:1 mixture 
showed numerous cracks while the 5:1 mixture showed better 
resistance to the heat treatment. When the 5:1 mixture was 
heated with 10 °C/h, the refractive index dropped from the 
initial 1.52 to 1.34 while the thickness decrease was about 28%. 
In case of the 5:5 mixture heating did not destroy the film, but 
the refractive index decreased to 1.44 [92]. 

Zhou et al. [34] deposited mixtures consisting of TMA+1,4- 
diaminobenzene hybrid and AI2O3 with the hybrid:oxide ratios 
of 1:1, 1:2, 1:3 and 1:4. The 1:4 mixture showed improved 
stability in ambient air: The film thickness decreased less than 
5% when kept at ambient air for several tens of weeks. The 
mixtures showed excellent electrical insulating properties. A 
current density of 2.3 x 10~ 8 A-cm -2 at an electric field of 
1 MV-cm"', and a dielectric constant of about 6.2 were 
measured for the 1:4 mixture. The 1:4 mixture showed clear 
charge trapping behavior, but not good enough to be used as a 
charge trap layer for a charge trap flash memory. 

Miller et al. [81] investigated the mechanical properties of 
AI2O3/TMA+EG/AI2O3 nanolaminates grown at 155 °C on 
polyethylene naphtalate substrates. The layer thicknesses were 
10/3/10 nm, 25/15/25 nm (shown in Figure 12), 25/3/25 nm, 
and 25/192/25 nm. The nanolaminates exhibited reduced crit- 
ical strains at fracture when compared to pure components. This 
was attributed to the low toughness of the TMA+EG alucone 
[81]. According to the microcantilever-facilitated curvature 
studies done later the curvature for the pure TMA+EG hybrid 
evolved when the films were stored in ambient air for two 



weeks. Investigation on 25/192/25 nm thick A1 2 0 3 /TMA+EG/ 
AI2O3 nanolaminates showed that the topmost AI2O3 layer 
stabilized the structure, possibly by shielding the underlying 
TMA+EG layer from moisture [84]. 



Nanolaminates fabricated using TMA+EG and AI2O3 were also 
investigated by Vaha-Nissi et al. [88]. The depositions were 
carried out at 100 °C on biopolymer (biaxially oriented poly- 
lactic acid) substrates, and the samples were characterized for 
their oxygen transmission rate (OTR) and water vapor transmis- 
sion rate (WVTR). It was shown that the five-layer nanolami- 
nates investigated, i.e., AI2O3/TMA+EG/AI2O3/TMA+EG/ 
AI2O3, worked essentially better than pure AI2O3 films as gas- 
barrier coatings on biopolymer materials. Purely inorganic coat- 
ings are somewhat brittle (as clearly evidenced from SEM 
images revealing large cracks for pure AI2O3 films on top of 
flexible biopolymer substrates), and straining them leads to 
defects and deteriorated gas-barrier properties. Apparently the 
intervening hybrid layers improve the flexibility of the coating. 
For the laminates the defect concentration was found to be 
considerably smaller compared to AI2O3 films after mechanical 
straining. The best barrier properties were achieved for the five- 
layer laminate deposited with 50 cycles of both components. 
The laminates were also found to be stable in air which was not 
the case for a thin AI2O3 layer alone on the biopolymer sub- 
strate. 

The WVTR of TMA+EG and AI2O3 coatings was investigated 
also by Park et al. [91]. The films were deposited at 85 °C on 
polyethylene naphthlate substrates. The WVTR values 
measured at 85 °C and 85% relative humidity for alucone, 
AI2O3 and Al 2 C>3/alucone nanolaminate coatings were ca. 1.1, 
0.037 and 0.021 g/(m 2 -day), respectively. The improved WVTR 
value of the nanolaminate coating when compared to the pure 
alucone and oxide layers alone was attributed to a synergy 




Pt/Au 
ALDAI 2 0 3 
MLD Alucone 

ALDAI2O3 
PEN 



100 nm 



Figure 12: A nanolaminate coating consisting of AI2O3 and TMA+EG 
alucone layers with targeted thicknesses of 25 and 15 nm, respective- 
ly, on a polyethylene naphtalate (PEN) substrate (reprinted with 
permission from [81], Copyright (2009) American Institute of Physics). 



1130 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



effect: As the alucone layer reacts readily with water, it reduces 
the diffusion speed as well as increases the diffusion paths of 
the water vapor. 

The Zn-based hybrid DEZ+HQ has been combined with ZnO to 
form mixtures which were anticipated to show enhanced elec- 
trical conductivity due to electron band overlap between the 
Jt-electrons of the HQ ring and ZnO [98,99]. The pure ZnO film 
had a conductivity of 14 S/cm in ambient light, whereas the 
pure hybrid film was nonconductive. The 1:1, 1:3 and 2:2 
mixtures of ZnO:DEZ+HQ had higher electrical conductance 
than ZnO, i.e., 1 16, 40 and 170 S/cm, respectively, whereas the 
1:5 and 5:5 mixtures exhibited lower conductance than ZnO, 
i.e., 6 and 13 S/cm. respectively. The elastic modulus varied 
depending on the precise composition of the mixtures, being 
about 30 GPa for a pure DEZ+HQ hybrid film, about 50 GPa 
for a 1:5 mixture, about 120 GPa for a 1:1 mixture, and about 
150 GPa for a pure ZnO film. Also the hardness of the films 
was affected by the moiety concentration, being about 1,4,5 
and 13 GPa for the same films, respectively. Pure DEZ+HQ, the 
1 : 1 mixture and the pure ZnO were all highly transparent in 
visible spectrum [99]. 

Superlattice structures have been successfully fabricated where 
nondoped or Al-doped ZnO layers of nanometer-scale alternate 
with extremely thin organic (hydroquinone) layers. The films 
were deposited with the DEZ+H 2 0, TMA+H 2 0 and DEZ+HQ 
processes at 220 °C, with the oxide:hybrid deposition cycle 
ratio varying from 199:1 to 1:1 [94,102,103]. FTIR and XRR 
studies confirmed the incorporation of organic layers and the 
formation of superlattice structures, respectively. When heated 
for one hour up to 700 °C, no changes in XRR patterns of the 
(Zn,Al)0:HQ films were observed until 450 °C, after which 
coarsening of the film started to inflict noise. The superlattice- 
originated XRR peaks were still present when annealed up to 
650 °C. No visual change in the films was observed until heat- 
treated at 700 °C. Apparently the ZnO layers protect the under- 
lying organic layers from decomposition at elevated tempera- 
tures. The eventual goal of the study was to suppress lattice 
thermal conductivity and/or enhance the Seebeck coefficient of 
ZnO films through the introduction of intervening organic 
layers such that the superlattice films could show overall 
enhanced thermoelectric characteristics, i.e., concomitant large 
Seebeck coefficient, high electrical conductivity and low 
thermal conductivity. Preliminary characterization of the 
(Zn,Al)0:HQ superlattice films confirmed that the films indeed 
showed promise as thermoelectric materials [94]. More recently 
it was demonstrated that similar superlattice structures 
consisting of thicker layers of ZnO combined with individual 
organic layers could be made not only with HQ but also with an 
AP or ODA layer [103]. The ZnO:hybrid ratio varied between 



199:1 and 39:1, and the superlattice structure was confirmed by 
XRR. Resistivity and Seebeck coefficient measurements 
showed an increase in carrier concentration for small concentra- 
tions of organic layer, whereas at higher concentrations a large 
reduction in carrier concentration was observed. 

Liu et al. [101] fabricated oxide-hybrid mixtures by using the 
same DEZ+H2O and DEZ+HQ processes as described above 
for the ZnO:HQ superlattices, and determined the thermal 
conductivity for the 1:1 mixture at about 0.13-0.15 W/(m-K). 
This value is much lower than what was measured by the same 
group for pure DEZ+HQ hybrid and discussed earlier in this 
review. It was suggested, that when employing structurally 
vastly different oxide and hybrid constituents in the material 
fabrication, the ZnO flakes and hybrid chains scatter efficiently 
phonons, resulting in a reduced thermal conductivity. The 
volumetric heat capacity for the 1:1 mixture was about 
2.9 J/(cm 3 -K), being only slightly less than what was reported 
for the DEZ+HQ system. 

The DEZ+AP and DEZ+H2O processes have been combined to 
form mixtures and nanolaminates. The crystallinity and density 
of the mixtures were varied by the number of hybrid and oxide 
cycles during the depositions. Although the hybrid containing 
films were unstable in air, which made the AFM measurements 
somewhat inaccurate, adding amorphous hybrid to crystalline 
ZnO had a smoothing effect on the samples. The nanolaminates 
which had a minimum of 3 nm thick ZnO layer on top were 
stable in ambient air and had a rather constant RMS roughness 
of 1-1.4 nm. According to the nanointendation measurements 
the DEZ+AP hybrid was soft, with a contact modulus of a 
typical polymer. Although ZnO is also soft, it is still harder and 
stiffer than the pure hybrid. Adding ZnO to the hybrid was 
observed to have little effect up to 1:1 hybrid:oxide ratio. From 
the nanointendation measurements performed on nanolami- 
nates it was concluded that the thicker the ZnO layer is, the 
more it enhances the hardness of the film and a thin ZnO layer 
can actually reduce film strength for some unknown reason. 
Wet-etching tests showed that adding ZnO does now improve 
the chemical stability of the mixtures. However, acetone treat- 
ment was observed to remove only the organic part of the film, 
leaving a porous oxide backbone [106]. 

Mixtures of Ti02 and TiCl4+ODA have been deposited with 
oxide:hybrid ratios from 1:1 to 10:1. A good control over the 
RMS roughness value and refractive index was achieved by 
varying the mixture composition. Also the density, reduced 
modulus, hardness and crystallinity of the material systemati- 
cally depended on the oxide:hybrid ratio. Wet-etching tests 
carried out with several solvents indicated that the mixtures 
were chemically extremely stable [113]. 



1131 



Beilstein J. Nanotechnol. 2014, 5, 1104-1136. 



For ZrC>2:ZTB+EG mixtures in which the oxide:hybrid ratio 
was varied from 1 :3 to 3: 1 [54,87], it was found that the refrac- 
tive index and elastic modulus changed continuously from 
values for pure ZTB+EG in ~ 1 .63 and E~21 GPa) to those of 
pure ZrC>2 (n ~ 1.86 and E ~ 97 GPa). Also density and hard- 
ness varied according to the oxide:hybrid ratio. 

Superlattice structures consisting of TiC>2 and T1CI4+HDD were 
successfully deposited at 100 °C by repetition of 50 cycles of 
Ti02 and one cycle of T1CI4+HDD. TEM images showed indi- 
vidual oxide and hybrid nanolayers. The films remained stable 
when annealed up to 400 °C [55]. 

As 7-OTS is a heterobifunctional material, for which the 
terminal C=C group is required to be converted to a carboxylic 
group by ozone treatment during the growth process, the 7-OTS 
hybrid materials grow in extremely well controlled layer-by- 
layer manner. Hybrid films based on 7-OTS have been used in 
several studies to fabricate various superlattice and nanolami- 
nate structures. Four bilayers consisting of 1.1 nm thick 
7-OTS+Ti(OCH(CH 3 ) 2 ) 4 hybrid and 2 nm thick Ti0 2 were 
used to fabricate nanolaminates. The nanolaminate was then 
sandwiched between Al metal wires to study the electrical prop- 
erties of the hybrid-oxide material. The constructed device had 
a large endurance and a long retention time, which demon- 
strated the potential application of the nanolaminates as 
nonvolatile memory materials [10]. 

Investigations on multilayered structures of 
7-OTS+Ti(OCH(CH 3 ) 2 ) 4 and Ti0 2 deposited on PEN indi- 
cated that the Ti02 blocked water permeation. The WVTR 
experiments suggested that the lag time for the structures was 
extended due to the tortuous path effect and water accumula- 
tion in the organic layers. For a structure consisting of five 
bilayers a WVTR value of 7.0 x 10~ 4 g/(m 2 -day) during a lag 
time of 155 h at 60 °C and a relative humidity 85% was 
obtained [40]. 

Superlattices with hybrid-Al203-hybrid-Ti02 structures with 
various mixing ratios have been fabricated using 
7-OTS+Ti(OCH(CH 3 ) 2 ) 4 and 7-OTS+TMA. The formed struc- 
tures when annealed were stable up to about 500 °C. The coat- 
ings showed good flexibility, were mechanically stable, and had 
various unique electrical properties. Organic pentacene thin- 
film transistors fabricated by using the superlattices on flexible 
plastic substrate had a drain current of 1.5 uA, a field effect 
mobility of 0.54 cm 2 /(V-s), and an inverter voltage gain 
-dF 0Ut /dKj n ~ 4.5 when operated at a voltage of -2 V [41]. 

A non-volatile flash memory thin-film transistor was made 
using 7-OTS+Ti(OCH(CH 3 ) 2 ) 4 and 7-OTS+TMA layers 



between ZnO and pentacene. The device showed promising 
non-volatile memory effects when operated at low voltages 
[42]. Organic pentacene thin- film transistors were also fabri- 
cated by using 7-OTS+ZTB and Zr02- A maximum field effect 
mobility of 0.63 cm 2 /(V-s) was measured, when operating at 
-1 V with an on/off current ratio of about 10 3 [43]. 

Nanolaminates consisting of 7-OTS+DEZ and ZnO layers have 
also been deposited. The thin-film transistors made by using the 
nanolaminate had a high field effect mobility of 7 cm 2 /(V-s), 
when operating at 3 V with an on/off current ratio of 10 6 and 
with a threshold voltage of 0.6 V. It was also concluded that the 
7-OTS+DEZ provides structural flexibility in the superlattice 
[44]. 

Han et al. [46] fabricated floating-gate nonvolatile memory 
transistors from two types of hybrid layers: Al-containing 
hybrid layers were deposited by using 7-OTS, water, O3 and 
TMA as precursors, whereas DEZ and HDD were used for the 
Zn-containing layers. Capacitor memory devices constructed by 
using Al-containing hybrid as blocking and tunneling layers 
with ZnO:Cu charge trap layer sandwiched between them 
(Figure 13), had a large memory window of 14.1 V operated at 
±15 V. The same structure was then used together with a 
Zn-containing hybrid as a semiconducting layer to form 
nonvolatile memory transistors which operated in voltage range 
of-1 to 3 V. The high writing/erasing (+8 V/— 12 V) current 
ratio of 10 3 obtained with the device indicated that the tested 
construction showed promise for memory electronics applica- 
tions. 




Figure 13: An HRTEM image of a capacitor memory device fabricated 
by using Al-containing hybrid (marked as AlOx-SAOL) as blocking and 
tunneling layers with a ZnO:Cu charge trap layer in between (reprinted 
with permission from [46], Copyright (2012) The Royal Society of 
Chemistry). 



1132 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



Conclusion 

The application perspectives of the original ALD thin-film tech- 
nology have become considerably wider through the introduc- 
tion of purely organic moieties as building units in the chem- 
ical atomic-scale controlled deposition process. Now by taking 
advantage of the MLD technique we are not only able to materi- 
alize in a molecular layer-by-layer manner high-quality thin 
films of various commercially attractive organic polymers but 
also of inorganic-organic hybrid materials, potentially 
combining the best attributes of the two entirely different 
chemistries. 

The organic polymers made up till now by means of the MLD 
technique include various amides, imides, imide-amides, ureas, 
urethanes, esters and imines, while in the case of the ALD/ 
MLD-grown inorganic-organic hybrid thin films the metal 
species variety covers the elements Al, Zn, Ti, Zr, Hf, V, Si, Fe 
and Mn. Although a rapidly increasing number of different 
precursors have already been exploited, the field is nothing but 
just approaching its emergent stage. Nevertheless essentially 
ideal MLD and ALD/MLD processes have already been devel- 
oped for several precursor combinations such that the film 
growth rates achieved well correspond to the values calculated 
on the bases of the expected lengths for straight polymer chains. 
Examples of such ideally behaving processes are the purely 
organic hexanedioyl dichloride+hexane-l,6-diamine and 
heptane- 1 ,7-diamine+nonanedioyl dichloride systems, and the 
hybrid diethylzinc+hexa-2,4-diyne-l,6-diol, TiCi4+hexa-2,4- 
diyne-l,6-diol and TiCl4+4-aminophenol systems. 

The layer-by- layer manner in which the films are grown in both 
ALD and MLD, provides us yet another powerful means of 
fine-tuning the film properties by depositing on-demand 
designed thin-film mixtures, superstructures and nanolaminates. 
Optical and mechanical properties, surface roughness and 
degree of crystallinity have been successfully tuned by mixing 
different deposition cycles, whereas control over the chemical 
stability, electrical and gas-barrier properties and electrical and 
thermal conductivities has been achieved by constructing well- 
defined superlattice and nanolaminate structures. Post-deposi- 
tion treatments of films containing organic moieties have also 
proven to further expand the application range of the ALD/ 
MLD fabricated hybrid thin films as such treatments enable, 
e.g., to produce porous coatings. 

The work on the ALD/MLD grown organic and 
inorganic-organic thin films is still in its beginning phase. 
Deeper studies are definitely required to shed light even on the 
precise growth mechanisms of these fundamentally new types 
of thin films, hopefully giving us better insight to select new 
well-behaving precursor pairs. As the huge potential of the 



hybrid films has been recognized, the number of articles 
featuring properties related to specific applications keeps rising. 
Recently, it was for example demonstrated that periodically 
repeating organic layers embedded in thicker inorganic layers 
can efficiently block heat conduction. This result is highly 
promising in the field of thermoelectrics. Tunable refractive 
index should on the other hand be extremely important for 
optical applications. As another example, nanolaminate struc- 
tures from oxides and hybrids improve the gas barrier prop- 
erties of the protective coatings. Especially noteworthy are also 
the porous structures, which could be used in optics, electronics 
and catalysis, to name just a few examples. In short, the 
prospects of the ALD/MLD fabricated films are excellent. 

Acknowledgements 

The present work has received funding from the European 
Research Council under the European Union's Seventh Frame- 
work Programme (FP/2007-2013)/ERC Advanced Grant Agree- 
ment (No. 339478), and also from the Academy of Finland (No. 
255562). Some of the early papers referred to here were found 
via the Virtual project on the history of ALD (VPHA). 

References 

1 . Sveshnikova, G. V.; Kol'tsov, S. I.; Aleskovskii, V. B. 
J. Appl. Chem. USSR 1970, 43, 432-434. 

2. Aleskovskii, V. B. J. Appl. Chem. USSR 1974, 47, 2145-2157. 

3. Suntola, T.; Antson, J. Method for producing compound thin films. 
U.S. Pat. Appl. US4058430 A, Nov 15, 1977. 

4. Puurunen, R. L. J. Appl. Phys. 2005, 97, 121301. 
doi:10.1063/1. 1940727 

5. Yoshimura, T.; Tatsuura, S.; Sotoyama, W. Appl. Phys. Lett. 1991, 59, 
482-484. doi:1 0.1 063/1. 105415 

6. Kubono, A.; Yuasa, N.; Shao, H.-L; Umemoto, S.; Okui, N. 
Thin Solid Films 1 996, 289, 1 07-1 1 1 . 

doi:1 0. 1 01 6/S0040-6090(96)0891 3-4 

7. Nagai, A.; Shao, H.; Umemoto, S.; Kikutani, T.; Okui, N. 
High Perform. Polym. 2001, 13, S169-S179. 

doi:1 0. 1 088/0954-0083/1 3/2/31 5 

8. Shao, H.; Umemoto, S.; Kikutani, T.; Okui, N. Polymer 1997, 38, 
459-462. doi: 1 0. 1 01 6/S0032-3861 (96)00504-6 

9. Yoshimura, T.; Tatsuura, S.; Sotoyama, W.; Matsuura, A.; Hayano, T. 
Appl. Phys. Lett. 1992, 60, 268-270. doi:10.1063/1. 106681 

10. Lee, B. H.; Ryu, M. K.; Choi, S.-Y.; Lee, K.-H.; Im, S.; Sung, M. M. 

J. Am. Chem. Soc. 2007, 129, 16034-16041. doi:10.1021/ja075664o 

11. Smirnov, V. M.; Zemtsova, E. G.; Belikov, A. A.; Zheldakov, I. L.; 
Morozov, P. E.; Polyachonok, O. G.; Aleskovskii, V. B. 

Dokl. Phys. Chem. 2007, 413, 95-98. 
doi : 1 0. 1 1 34/S00 1 250 1 607040069 

12. Dameron, A. A.; Seghete, D.; Burton, B. B.; Davidson, S. D.; 
Cavanagh, A. S.; Bertrand, J. A.; George, S. M. Chem. Mater. 2008, 
20, 3315-3326. doi:10.1021/cm7032977 

13. Nilsen, O.; Klepper, K. B.; Nielsen, H. 0.; Fjellvag, H. ECS Trans. 
2008, 16, 3-14. doi:10.1 149/1 .2979975 

14. Sood, A.; Sundberg, P.; Malm, J.; Karppinen, M. Appl. Surf. Sci. 2011, 
257, 6435-6439. doi:10.1016/j.apsusc.201 1 .02.022 



1133 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



15. Liang, X.; Yu, M.; Li, J.; Jiang, Y.-B.; Weimer, A. W. Chem. Commun. 

2009, 7140-7142. doi:10.1039/b911888h 

16. Seghete, D.; Davidson, B. D.; Hall, R. A.; Chang, Y. J.; Bright, V. M.; 
George, S. M. Sens. Actuators, A 2009, 755, 8-15. 
doi:10.1016/j.sna.2008.12.016 

17. Gabriel, N. T.; Talghader, J. J. J. Appl. Phys. 2011, 770, 043526. 
doi:10.1063/1.3626462 

18. Smith, S. W.; McAuliffe, K. G.; Conley, J. F., Jr. Solid-State Electron. 

2010, 54, 1076-1082. doi:10.1016/j.sse.2010.05.007 

19. Heo, J.; Liu, Y.; Sinsermsuksakul, P.; Li, Z.; Sun, L; Noh, W.; 
Gordon, R. G. J. Phys. Chem. C2011, 775, 10277-10283. 
doi:10.1021/jp202202x 

20. Elam, J. W.; Sechrist, Z. A.; George, S. M. Thin Solid Films 2002, 474, 
43-55. doi:10.1016/S0040-6090(02)00427-3 

21. George, S. M. Chem. Rev. 2010, 770, 111-131. 
doi:10.1021/cr900056b 

22. Miikkulainen, V.; Leskela, M.; Ritala, M.; Puurunen, R. L. 
J. Appl. Phys. 2013, 773, 021301. doi:10.1063/1.4757907 

23. Knez, M.; Nielsch, K.; Niinisto, L. Adv. Mater. 2007, 79, 3425-3438. 
doi:10.1002/adma.200700079 

24. George, S. M.; Yoon, B.; Dameron, A. A. Acc. Chem. Res. 2009, 42, 
498-508. doi:10.1021/ar800105q 

25. Leskela, M.; Ritala, M.; Nilsen, O. MRS Bull. 2011, 36, 877-884. 
doi:10.1557/mrs.201 1.240 

26. George, S. M. The Strem Chemiker 201 1 , XXV, 13-26. 

27. George, S. M.; Lee, B. H.; Yoon, B.; Abdulagatov, A. I.; Hall, R. A. 
J. Nanosci. Nanotechnol. 2011, 77, 7948-7955. 
doi:10.1166/jnn.201 1.5034 

28. Lee, B. H.; Yoon, B.; Abdulagatov, A. I.; Hall, R. A.; George, S. M. 
Adv. Fund. Mater. 2013, 23, 532-546. doi:10.1002/adfm.201200370 

29. Yoshimura, T.; Yoshino, C; Sasaki, K.; Sato, T.; Seki, M. 
IEEE J. Sel. Top. Quantum Electron. 2012, 78, 1192-1199. 
doi:1 0. 1 1 09/JSTQE.201 1 .21 67676 

30. King, D. M.; Liang, X.; Weimer, A. W. ECS Trans. 2009, 25, 163-190. 
doi:10.1 149/1 .3205053 

31. Zhou, H.; Bent, S. F. J. Vac. Sci. Technol., A 2013, 37, 040801. 
doi:10.1 116/1 .4804609 

32. Klepper, K. B.; Nilsen, O.; Fjellvag, H. Dalton Trans. 2010, 39, 
1 1 628-1 1 635. doi: 1 0.1 039/c0dt0081 7f 

33. Yoon, B.; Lee, Y.; Derk, A.; Musgrave, C; George, S. M. ECS Trans. 

2011, 33, 191-195. doi:10.1 149/1 .3565514 

34. Zhou, W.; Leem, J.; Park, I.; Li, Y.; Jin, Z.; Min, Y.-S. J. Mater. Chem. 

2012, 22, 23935-23943. doi:10.1039/c2jm35553a 

35. Nilsen, O.; Haug, K. R.; Finstad, T.; Fjellvag, H. 
Chem. Vap. Deposition 2013, 79, 174-179. 
doi : 1 0. 1 002/cvde.20 1 207043 

36. Sood, A.; Sundberg, P.; Karppinen, M. Dalton Trans. 2013, 42, 
3869-3875. doi:10.1039/c2dt32630b 

37. Sundberg, P.; Karppinen, M. Eur. J. Inorg. Chem. 2014, 968-974. 
doi:10.1002/ejic.201301560 

38. Gong, B.; Peng, Q.; Parsons, G. N. J. Phys. Chem. B 2011, 775, 
5930-5938. doi:1 0. 1 021/jp2011 86k 

39. Lee, Y.; Yoon, B.; Cavanagh, A. S.; George, S. M. Langmuir 2011 , 27, 
1 51 55-1 5164. doi:10.1 021/la202391 h 

40. Seo, S.-W.; Jung, E.; Lim, C; Chae, H.; Cho, S. M. Thin Solid Films 
2012, 520, 6690-6694. doi:10.1016/j.tsf.2012.07.017 

41. Lee, B. H.; Lee, K. H.; Im, S.; Sung, M. M. Org. Electron. 2008, 9, 
1146-1153. doi:10.1016/j.orgel.2008.08.015 

42. Cha, S. H.; Park, A.; Lee, K. H.; Im, S.; Lee, B. H.; Sung, M. M. 
Org. Electron. 2010, 77, 159-163. doi:10.1016/j.orgel.2009.09.021 



43. Lee, B. H.; Im, K. K.; Lee, K. H.; Im, S.; Sung, M. M. Thin Solid Films 
2009, 577, 4056-4060. doi:10.1016/j.tsf.2009.01.173 

44. Park, Y.; Han, K. S.; Lee, B. H.; Cho, S.; Lee, K. H.; Im, S.; 
Sung, M. M. Org. Electron. 2011, 72, 348-352. 
doi:10.1016/j.orgel.2010.1 1.026 

45. Huang, J.; Lee, M.; Lucero, A.; Kim, J. Chem. Vap. Deposition 2013, 
79, 142-148. doi:10.1002/cvde.201207041 

46. Han, K. S.; Park, Y.; Han, G.; Lee, B. H.; Lee, K. H.; Son, D. H.; 
Im, S.; Sung, M. M. J. Mater. Chem. 2012, 22, 19007-19013. 
doi:10.1039/c2jm32767h 

47. Gong, B.; Parsons, G. N. ECS J. Solid State Sci. Technol. 2012, 7, 
P21 0-P21 5. doi: 1 0.1 1 49/2.023204jss 

48. Yoon, B.; Seghete, D.; Cavanagh, A. S.; George, S. M. Chem. Mater. 
2009, 27, 5365-5374. doi:10.1021/cm9013267 

49. Chen, C; Li, P.; Wang, G.; Yu, Y.; Duan, F.; Chen, C; Song, W.; 
Qin, Y.; Knez, M. Angew. Chem., Int. Ed. 2013, 52, 9196-9200. 
doi:10.1002/anie.201302329 

50. Du, Y.; George, S. M. J. Phys. Chem. C2007, 777, 8509-8517. 
doi:10.1021/jp067041n 

51. Peng, Q.; Gong, B.; Van Gundy, R. M.; Parsons, G. N. Chem. Mater. 
2009, 27, 820-830. doi:10.1021/cm8020403 

52. Yoon, B.; O'Patchen, J. L.; Seghete, D.; Cavanagh, A. S.; 
George, S. M. Chem. Vap. Deposition 2009, 75, 112-121. 
doi:10.1002/cvde.200806756 

53. Abdulagatov, A.; Hall, R. A.; Sutherland, J. L.; Lee, B. H.; 
Cavanagh, A. S.; George, S. M. Chem. Mater. 2012, 24, 2854-2863. 
doi:10.1021/cm300162v 

54. Lee, B. H.; Anderson, V. R.; George, S. M. Chem. Vap. Deposition 
2013, 79, 204-212. doi:10.1002/cvde.201207045 

55. Cho, S.; Han, G.; Kim, K.; Sung, M. M. Angew. Chem., Int. Ed. 2011, 
50, 2742-2746. doi: 10.1002/anie.201 00631 1 

56. Yoon, K.-H.; Han, K.-S.; Sung, M.-M. Nanoscale Res. Lett. 2012, 7, 
71. doi:10.1186/1556-276X-7-71 

57. Klepper, K. B.; Nilsen, O.; Hansen, P.-A.; Fjellvag, H. Dalton Trans. 
2011, 40, 4636-4646. doi:10.1039/c0dt01716g 

58. Design Institute for Physical Properties, DIPPR Project 801 - Full 
Version, Design Institute for Physical Property Data/AIChE, 2005. 

59. Peng, Q.; Efimenko, K.; Genzer, J.; Parsons, G. N. Langmuir 2012, 
28, 10464-10470. doi:10.1021/la3017936 

60. Adamczyk, N. M.; Dameron, A. A.; George, S. M. Langmuir 200%, 24, 
2081-2089. doi:10.1021/la7025279 

61. Putkonen, M.; Harjuoja, J.; Sajavaara, T.; Niinisto, L. J. Mater. Chem. 
2007, 77, 664-669. doi:10.1039/b612823h 

62. Salmi, L. D.; Puukilainen, E.; Vehkamaki, M.; Heikkila, M.; Ritala, M. 
Chem. Vap. Deposition 2009, 75, 221-226. 
doi:10.1002/cvde.200906770 

63. Yoshida, S.; Ono, T.; Esashi, M. Micro Nano Lett. 2010, 5, 321-323. 
doi:10.1049/mnl.2010.0128 

64. Yoshida, S.; Ono, T.; Esashi, M. Nanotechnology 2011, 22, 335302. 
doi : 1 0. 1 088/0957-4484/22/33/335302 

65. Haq, S.; Richardson, N. V. J. Phys. Chem. B 1999, 77, 5256-5265. 
doi:10.1021/jp984813+ 

66. Miyamae, T.; Tsukagoshi, K.; Matsuoka, O.; Yamamoto, S.; 
Nozoye, H. Jpn. J. Appl. Phys., Part 1 2002, 47, 746-748. 
doi:10.1143/JJAP.41.746 

67. Bitzer, T.; Richardson, N. V. Appl. Phys. Lett. 1997, 77, 662-664. 
doi:10.1063/1. 119822 

68. Bitzer, T.; Richardson, N. V. Appl. Surf. Sci. 1999, 144-145, 339-343. 
doi:1 0. 1 01 6/S01 69-4332(98)00823-X 



1134 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



69. Kim, A.; Filler, M. A.; Kim, S.; Bent, S. F. J. Am. Chem. Soc. 2005, 
127, 6123-6132. doi:10.1021/ja042751x 

70. Loscutoff, P. W.; Zhou, H.; Clendenning, S. B.; Bent, S. F. ACS Nano 

2010, 4, 331-341. doi:10.1021/nn901013r 

71. Prasittichai, C; Zhou, H.; Bent, S. F. ACS Appl. Mater. Interfaces 
2013, 5, 13391-13396. doi:10.1021/am4043195 

72. Zhou, H.; Bent, S. F. ACS Appl. Mater. Interfaces 2011 , 3, 505-511. 
doi:10.1021/am1010805 

73. Zhou, H.; Toney, M. F.; Bent, S. F. Macromolecules 2013, 46, 
5638-5643. doi:10.1021/ma400998m 

74. Lee, J. S.; Lee, Y.-J.; Tae, E. L; Park, Y. S.; Yoon, K. B. Science 
2003, 301, 818-821. doi:10.1126/science.1086441 

75. Loscutoff, P. W.; Lee, H.-B.-R.; Bent, S. F. Chem. Mater. 2010, 22, 
5563-5569. doi:10.1021/cm1016239 

76. Ivanova, T. V.; Maydannik, P. S.; Cameron, D. C. 

J. Vac. Sci. Technol., A 2012, 30, 01 A121 . doi:10.1 1 16/1 .3662846 

77. Yoshimura, T.; Kudo, Y. Appl. Phys. Express 2009, 2, 015502. 
doi:10.1143/APEX.2.015502 

78. Yoshimura, T.; Ito, S.; Nakayma, T.; Matsumoto, K. Appl. Phys. Lett. 
2007, 91, 033103. doi:10.1063/1 .2754646 

79. Yoshimura, T.; Ebihara, R.; Oshima, A. J. Vac. Sci. Technol., A 2011, 
29, 051510. doi:10.1116/1 .3620644 

80. Yoshimura, T.; Ishii, S. J. Vac. Sci. Technol., A 2013, 31, 031501. 
doi:10.1116/1.4793478 

81. Miller, D. C; Foster, R. R.; Zhang, Y.; Jen, S.-H.; Bertrand, J. A.; 
Lu, Z.; Seghete, D.; O'Patchen, J. L.; Yang, R.; Lee, Y.-C; 
George, S. M.; Dunn, M. L. J. Appl. Phys. 2009, 105, 093527. 
doi:10.1063/1 .3124642 

82. Liang, X.; King, D. M.; Li, P.; George, S. M.; Weimer, A. W.AIChEJ. 
2009, 55, 1030-1039. doi:10.1002/aic.11757 

83. Qin, Y.; Yang, Y.; Scholz, R.; Pippel, E.; Lu, X.; Knez, M. Nano Lett. 

2011, 11, 2503-2509. doi:10.1021/nl2010274 

84. Miller, D. C; Foster, R. R.; Jen, S.-H.; Bertrand, J. A.; Seghete, D.; 
Yoon, B.; Lee, Y.-C; George, S. M.; Dunn, M. L. Acta Mater. 2009, 
57, 5083-5092. doi:10.1016/j.actamat.2009.07.015 

85. Liang, X.; Weimer, A. W. J. Nanopart. Res. 2010, 12, 135-142. 
doi:1 0. 1 007/sl 1 051-009-9587-0 

86. Lee, B. H.; Yoon, B.; Anderson, V. R.; George, S. M. 

J. Phys. Chem. C2012, 116, 3250-3257. doi:10.1021/jp209003h 

87. Lee, B. H.; Anderson, V. R.; George, S. M. ECS Trans. 2011, 41, 
131-138. doi:10.1 149/1 .3633661 

88. Vaha-Nissi, M.; Sundberg, P.; Kauppi, E.; Hirvikorpi, T.; Sievanen, J.; 
Sood, A.; Karppinen, M.; Harlin, A. Thin Solid Films 2012, 520, 
6780-6785. doi:10.1016/j.tsf.2012.07.025 

89. Yu, M.; Funke, H. H.; Noble, R. D.; Falconer, J. L. J. Am. Chem. Soc. 
2011, 133, 1748-1750. doi:10.1021/ja108681n 

90. Oldham, C. J.; Gong, B.; Spagnola, J. C; Jur, J. S.; Senecal, K. J.; 
Godfrey, T. A.; Parsons, G. N. J. Electrochem. Soc. 2011, 158, 
D549-D556. doi:10.1 149/1 .3609046 

91. Park, M.; Oh, S.; Kim, H.; Jung, D.; Choi, D.; Park, J.-S. 

Thin Solid Films 2013, 546, 153-156. doi:1 0. 1 016/j.tsf.201 3.05.01 7 

92. Ghazaryan, L.; Kley, E.-B.; Tunnermann, A.; Szeghalmi, A. V. 

J. Vac. Sci. Technol., A 2013, 31, 01A149. doi:10.1 1 16/1 .4773296 

93. Oldham, C. J.; Gong, B.; Spagnola, J.; Jur, J.; Senecal, K. J.; 
Godfrey, T. A.; Parsons, G. N. ECS Trans. 2010, 33, 279-290. 
doi:10.1149/1.3485265 

94. Tynell, T.; Terasaki, I.; Yamauchi, H.; Karppinen, M. 

J. Mater. Chem. A 2013, 1, 13619-13624. doi:10.1039/c3ta12909h 

95. Brown, J. J.; Hall, R. A.; Kladitis, P. E.; George, S. M.; Bright, V. M. 
ACS Nano 2013, 7, 7812-7823. doi:10.1021/nn402733g 



96. Jen, S.-H.; George, S. M.; McLean, R. S.; Carcia, P. F. 
ACS Appl. Mater. Interfaces 201 3, 5, 1165-1173. 
doi:10.1021/am303077x 

97. Gong, B.; Spagnola, J. C; Parsons, G. N. J. Vac. Sci. Technol., A 
2012, 30, 01A156. doi:10.1 1 16/1 .3670963 

98. Liang, X.; Jiang, Y.-B.; Weimer, A. W. J. Vac. Sci. Technol., A 2012, 
30, 01A108. doi:10.1 1 16/1 .3644952 

99. Yoon, B.; Lee, B. H.; George, S. M. ECS Trans. 2011, 41, 271-277. 
doi:10.1 149/1 .3633677 

100. Yoon, B.; Lee, B. H.; George, S. M. J. Phys. Chem. C2012, 116, 
24784-2479 1 . doi: 1 0. 1 02 1 /jp3057477 

101. Liu, J.; Yoon, B.; Kuhlmann, E.; Tian, M.; Zhu, J.; George, S. M.; 
Lee, Y.-C; Yang, R. Nano Lett. 2013, 13, 5594-5599. 
doi:10.1021/nl403244s 

102. Tynell, T.; Karppinen, M. Thin Solid Films 2014, 551, 23-26. 
doi:10.1016/j.tsf.2013.11.067 

103. Tynell, T.; Yamauchi, H.; Karppinen, M. J. Vac. Sci. Technol., A 2014, 
32, 01A105. doi: 1 0.1 11 6/1 .4831 751 

104.lshchuk, S.; Taffa, D. H.; Hazut, O.; Kaynan, N.; Yerushalmi, R. 

ACS Nano 2012, 6, 7263-7269. doi:10.1021/nn302370y 
105.Abdulagatov, A. I.; Terauds, K. E.; Travis, J. J.; Cavanagh, A. S.; 

Raj, R.; George, S. M. J. Phys. Chem. C2013, 117, 17442-17450. 

doi:10.1021/jp4051947 
106. Sundberg, P.; Sood, A.; Liu, X.; Karppinen, M. Dalton Trans. 2013, 42, 

1 5043-1 5052. doi: 1 0.1 039/c3dt51 578h 
107.Bahlawane, N.; Arl, D.; Thomann, J.-S.; Adjeroud, N.; Menguelti, K.; 

Lenoble, D. Surf. Coat. Technol. 2013, 230, 101-105. 

doi:10.1016/j.surfcoat.201 3.06. 098 
108. Seghete, D.; Hall, R. A.; Yoon, B.; George, S. M. Langmuir 201 0, 26, 

19045-19051. doi:10.1021/la102649x 
109. Liang, X.; Evanko, B. W.; Izar, A.; King, D. M.; Jiang, Y.-B.; 

Weimer, A. W. Microporous Mesoporous Mater. 2013, 168, 178-182. 

doi:10.1016/j.micromeso.2012. 09.035 
HO.KIepper, K. B.; Nilsen, O.; Levy, T.; Fjellvag, H. Eur. J. Inorg. Chem. 

2011, 2011, 5305-5312. doi:10.1002/ejic.201100192 
111. Salmi, L. D.; Heikkila, M. J.; Puukilainen, E.; Sajavaara, T.; 

Grosso, D.; Ritala, M. Microporous Mesoporous Mater. 2013, 182, 

147-154. doi:10.1016/j.micromeso.2013.08.024 
112.Klepper, K. B.; Nilsen, O.; Francis, S.; Fjellvag, H. Dalton Trans. 2014, 

43, 3492-3500. doi:10.1039/c3dt52391h 
113. Sundberg, P.; Sood, A.; Liu, X.; Johansson, L.-S.; Karppinen, M. 

Dalton Trans. 2012, 41, 10731-10739. doi:10.1039/c2dt31026k 

114. Kao, C.-Y.; Yoo, J.-W.; Min, Y.; Epstein, A. J. 
ACS Appl. Mater. Interfaces 201 2, 4, 137-141. 
doi:10.1021/am201506h 

115. Zhou, H.; Bent, S. F. J. Phys. Chem. C2013, 117, 19967-19973. 
doi:10.1021/jp4058725 

116.Smirnov, V.; Zemtsova, E.; Morozov, P. Rev. Adv. Mater. Sci. 2009, 
21, 205-210. 



1135 



BeilsteinJ. Nanotechnol. 2014, 5, 1104-1136. 



License and Terms 

This is an Open Access article under the terms of the 
Creative Commons Attribution License 
( http://creativecommons.Org/licenses/by/2.0 ), which 
permits unrestricted use, distribution, and reproduction in 
any medium, provided the original work is properly cited. 

The license is subject to the Beilstein Journal of 
Nanotechnology terms and conditions: 
( http ://www.beilstein- journals .org/b j nano ) 

The definitive version of this article is the electronic one 

which can be found at: 

doi:10.3762/bjnano.5.123 



1136